+ All Categories
Home > Documents > [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant...

[Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant...

Date post: 09-Dec-2016
Category:
Upload: kees
View: 213 times
Download: 1 times
Share this document with a friend
34
Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes Gerd Patrick Bienert and Franc ¸ois Chaumont Abstract Exchange across biological membranes is controlled by the composition of the lipid bilayer, diffusion-facilitating channels, and active transport proteins. In 1992, a protein facilitating the passive diffusion of water across membranes was discovered in humans and named aquaporin-1. Since then, an increasing number of proteins belonging to the same superfamily of membrane intrinsic proteins have been identified and characterized as ubiquitous indispensable players in transmem- brane water fluxes and water homeostasis. Compared to all other kingdoms of life, plants possess a high number of isoforms, clustered into seven subfamilies. A fascinating diversity of small, water-soluble, and uncharged compounds, ranging from gases to metalloids, has been identified as substrates for plant aquaporins. This chapter summarizes a variety of features and transport properties of these mem- brane pores illustrating their physiologically crucial contribution to water homeo- stasis, nutrition, and signaling processes. 1 Introduction Plant cells and their internal organelles are separated from their environment by biological membranes, generating independent, but highly controllable, functional units and microenvironments. However, the functionality of these units is strongly dependent on the controlled exchange of information and substances across these barriers. Biological membranes consist of a lipid bilayer with a highly hydrophobic interior, which restricts the diffusion of charged and polar molecules. Because of the nature of the membrane, it can be stated that the bigger and more polar a compound, the worse it will permeate through the lipid bilayer. Thus, only small G.P. Bienert and F. Chaumont (*) Institute of Life Sciences, Universite ´ catholique de Louvain, Croix du Sud, 4-15, 1348 Louvain-la- Neuve, Belgium e-mail: [email protected]; [email protected] M. Geisler and K. Venema (eds.), Transporters and Pumps in Plant Signaling, Signaling and Communication in Plants 7, DOI 10.1007/978-3-642-14369-4_1, # Springer-Verlag Berlin Heidelberg 2011 3
Transcript
Page 1: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Plant Aquaporins: Roles in Water Homeostasis,

Nutrition, and Signaling Processes

Gerd Patrick Bienert and Francois Chaumont

Abstract Exchange across biological membranes is controlled by the composition

of the lipid bilayer, diffusion-facilitating channels, and active transport proteins. In

1992, a protein facilitating the passive diffusion of water across membranes was

discovered in humans and named aquaporin-1. Since then, an increasing number of

proteins belonging to the same superfamily of membrane intrinsic proteins have

been identified and characterized as ubiquitous indispensable players in transmem-

brane water fluxes and water homeostasis. Compared to all other kingdoms of life,

plants possess a high number of isoforms, clustered into seven subfamilies. A

fascinating diversity of small, water-soluble, and uncharged compounds, ranging

from gases to metalloids, has been identified as substrates for plant aquaporins. This

chapter summarizes a variety of features and transport properties of these mem-

brane pores illustrating their physiologically crucial contribution to water homeo-

stasis, nutrition, and signaling processes.

1 Introduction

Plant cells and their internal organelles are separated from their environment by

biological membranes, generating independent, but highly controllable, functional

units and microenvironments. However, the functionality of these units is strongly

dependent on the controlled exchange of information and substances across these

barriers. Biological membranes consist of a lipid bilayer with a highly hydrophobic

interior, which restricts the diffusion of charged and polar molecules. Because of

the nature of the membrane, it can be stated that the bigger and more polar a

compound, the worse it will permeate through the lipid bilayer. Thus, only small

G.P. Bienert and F. Chaumont (*)

Institute of Life Sciences, Universite catholique de Louvain, Croix du Sud, 4-15, 1348 Louvain-la-

Neuve, Belgium

e-mail: [email protected]; [email protected]

M. Geisler and K. Venema (eds.), Transporters and Pumps in Plant Signaling,Signaling and Communication in Plants 7,

DOI 10.1007/978-3-642-14369-4_1, # Springer-Verlag Berlin Heidelberg 2011

3

Page 2: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

and nonpolar compounds can easily permeate unaided through biological mem-

branes. Regardless of the chemical nature of the compound, some are transported

across the membrane against the electrochemical gradient resulting in their accu-

mulation above the equilibrium concentration. The transport of compounds, espe-

cially that of larger, polar, and charged molecules, has therefore to be facilitated and

controlled by membrane proteins (see other chapters in this book).

For a long time, it was thought that small polar substances were transported

across biological membranes solely by passive diffusion and that there were no

protein-facilitated pathways. This was thought to be the case for the uncharged

water molecule, which is the most abundant substance in cells and organisms and is

indispensable for life. However, it soon became obvious that the nonfacilitated

passive diffusion of water across membranes was not in agreement with the

observed high membrane permeability of various cells (Agre et al. 1993). Addi-

tionally, the activation energy of water movement across some cell membranes was

found to be much lower than the values obtained for pure artificial membranes

(Ea > 10 kcal/mol). Similar discrepancies between in vivo and calculated or in vitro

measured permeability parameters were also observed for other small uncharged

molecules (Dordas et al. 2000). These observations could only be explained by

protein-mediated facilitated diffusion of these solutes. In 1956, Stein and Danielli

hypothesized that the high membrane water permeability of red blood cells must be

due to hydrophilic protein pores (Stein and Danielli 1956). Although their protein

nature remained undiscovered for a long time, the existence of such water channels

was demonstrated in 1992 by Preston in Agre’s laboratory, who functionally

characterized the first aquaporin (AQP) (CHIP28, renamed AQP1) from human

red blood cells (Preston et al. 1992), a major breakthrough in the field. It was soon

observed that AQP1 had sequence similarity to other known, but functionally

uncharacterized, proteins from mammals and plants. Intensive research concen-

trated on the identification of an increasing number of related AQP proteins in

different organisms and their characterization with respect to cell water homeosta-

sis. The crucial role of AQPs, at this time confined to water transport processes,

became clear, and the award of the Nobel Prize in Chemistry to Peter Agre in 2003

acknowledged their discovery and their relevance for life.

In the following years, a variety of important metabolic small uncharged solutes

were recognized as substrates for AQPs, and several of these channels were proven

to play important roles in the uptake, translocation, sequestration, or extrusion of

these solutes. AQPs have been identified in vertebrates, insects, plants, fungi,

protozoa, bacteria, and even viruses, thus covering all kingdoms of life (Zardoya

2005; Gazzarrini et al. 2006). Based on sequence homologies and in agreement with

their major transport activities, phylogenetic analyses gave rise to two clear clades:

the “water-permeable” AQPs, forming the AQP cluster, and the “glycerol-perme-

able” aquaglyceroporins, forming the glycerol facilitator-like protein (GLP) cluster.

While the aquaglyceroporins from prokaryotes and eukaryotes most likely com-

prise a monophyletic clade, the AQPs represent a heterogeneous group of all the

nonaquaglyceroporin isoforms (Danielson and Johanson 2010) Together, these two

clades constitute the superfamily of major intrinsic proteins (MIPs), which form

4 G.P. Bienert and F. Chaumont

Page 3: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

these fascinating hydrophilic pathways responsible for the passage of small

uncharged molecules through biological membranes. In all kingdoms of life,

MIPs play key roles in highly diverse and important physiological processes. In

microbes, the functions of MIPs range from osmoadaptation or turgor regulation to

enhanced cellular tolerance to rapid freezing and drug resistance (Petterson et al.

2005). Studies on mammalian AQP knock-out mice have pointed to the important

functions of MIP-facilitated water and glycerol transport in transepithelial fluid

transport, the influx and efflux of water in the brain, cell migration, sensory

signaling, the glycerol content of diverse tissues, skin hydration, cell proliferation,

carcinogenesis, and fat metabolism (reviewed in Verkman 2009).

2 Plant Aquaporins

One remarkable difference between MIPs from plants and other organisms is the

much larger isoform diversity in plants. While 13 isoforms have been identified in

mammals, many more have been found in the genomes of higher plants, with 35 in

Arabidopsis (Johanson et al. 2001), 33 in rice (Sakurai et al. 2005), 37 in tomato

(Sade et al. 2009), 45 in poplar (Gupta and Sankararamakrishnan 2009), and at least

36 in maize (Chaumont et al. 2001). Even the evolutionarily early land plant

Physcomitrella patens has 23 different isoforms (Danielson and Johanson 2008).

Plant MIPs can be subdivided into seven evolutionarily distinct subfamilies.

While the plasma membrane intrinsic proteins (PIPs), tonoplast intrinsic proteins

(TIPs), small basic intrinsic proteins (SIPs), nodulin26-like intrinsic proteins

(NIPs), hybrid intrinsic proteins (HIPs), and X intrinsic proteins (XIPs) all belong

to the AQP cluster of MIPs, the glycerol facilitator (GlpF)-like intrinsic proteins

(GIPs) represent the only known plant protein belonging to the GLP cluster. To

date, EST sequence data have revealed that all land plants, ranging from nonvascu-

lar plants to angiosperms, contain MIPs belonging to the PIP, TIP, NIP, and SIP

subfamilies. During the evolution of higher plants, the XIP subfamily was lost in

monocots and some dicot species. HIPs and GIPs have not yet been found in any

vascular plant (Fig. 1).

2.1 Plasma Membrane Intrinsic Proteins

PIPs constitute the most homogeneous group amongst the AQP subfamilies. The 15

isoforms found in poplar and the 13 found in each of Arabidopsis, rice, and maize

can be further subdivided into two groups, named PIP1 and PIP2. Although

Zea mays PIP1s (ZmPIP1s) have a longer N terminus, a shorter C terminus, and

a smaller extracellular loop A than ZmPIP2s. The overall sequence identity is

higher than 50% (Chaumont et al. 2001). PIPs are mainly localized to the plasma

membrane (Table 1). PIP2 proteins are highly efficient water channels when

Plant Aquaporins 5

Page 4: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

expressed in heterologous systems, while PIP1s are often inactive or possess a

much lower activity (reviewed in Chaumont et al. 2005; Kaldenhoff and Fischer

2006). A few isoforms have been shown to additionally facilitate the transport of

urea, hydrogen peroxide (H2O2), carbon dioxide (CO2), or boric acid (Table 2)

(Uehlein et al. 2003; Bienert et al. 2008b; Dynowski et al. 2008a, b; Uehlein et al.

2008; Fitzpatrick and Reid 2009). The expression of the different PIP isoforms is

quite high, and PIPs are expressed ubiquitously throughout the plant. Many factors

influence their expression (Table 3).

2.2 Tonoplast Intrinsic Proteins

The TIP subfamily derives its name from their main localization in the tonoplast.

TIPs can account for up to 40% of the total tonoplast proteins (Higuchi et al. 1998).

Different TIPs seem to be specific for different vacuoles, such as storage or lytic

Fig. 1 Phylogenetic classification of plant aquaporins. PIP, TIP, NIP, and SIP isoforms have been

identified in all plant species so far. Typical aquaporin sequences from Arabidopsis (At;

A. thaliana) were used to generate the phylogenetic tree (Johanson et al. 2001). Additionally,

aquaporin sequences of XIP, HIP, and GIP isoforms, which are absent in diverse plant species, for

example, Arabidopsis, were used from the moss Physcomitrella patents (Pp) (Danielson and

Johanson 2008)

6 G.P. Bienert and F. Chaumont

Page 5: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

vacuoles (Paris et al. 1996; Jauh et al. 1999; Park et al. 2004). However, this

specificity might be restricted to certain plants or developmental stages (Olbrich

et al. 2007). Mass spectrometry studies have shown that TIPs are also present in the

plasma membrane fraction (Table 1) (Jauh et al. 1999; Santoni et al. 2003; White-

man et al. 2008). In Arabidopsis, maize, rice, and poplar, 10, 13, 11, and 11

isoforms have been found, respectively, and they can be clustered into five phylo-

genetic groups, TIP1–TIP5 (Johanson et al. 2001). This phylogenetic division is

applicable to all vascular plants. TIPs possess a very high water channel activity in

all plant species (reviewed in Kaldenhoff and Fischer 2006). Additional substrate

specificities for urea, ammonia, and H2O2 have been observed when certain iso-

forms were heterologously expressed (Table 2) (Liu et al. 2003; Jahn et al. 2004;

Loque et al. 2005; Bienert et al. 2007; Dynowski et al. 2008a, b).

2.3 Small Basic Intrinsic Proteins

The SIP subfamily constitutes a very small, but the most divergent, AQP subfamily,

with 3–6 isoforms in Arabidopsis, maize, rice, or poplar. SIPs show very low

sequence identity with the other subfamilies and form two groups, which have

been localized to the endoplasmic reticulum. While AtSIP1 isoforms have been

shown to channel water across vesicle membranes (Ishikawa et al. 2005), no

substrate for SIP2 isoforms has yet been identified and their physiological role

remains obscure.

Table 1 Subcellular localization of some plant MIP isoforms

Subcellular localization MIP isoform References

Plasma membrane PIPs in general

AtNIP5;1 Takano et al. (2006)

OsNIP2;1 Ma et al. (2006)

AtNIP2;1 Choi and Roberts (2007)

TIPs Whiteman et al. (2008), Marmagne et al. (2004)

Tonoplast TIPs in general

PIPs Barkla et al. (1999), Shimaoka et al. (2004)

Endoplasmic reticulum AtSIPs Ishikawa et al. (2005)

AtNIP2;1 Mitzutani et al. (2006)

ZmPIP1s Zelazny et al. (2007)

Mitochondrion rAQP8a Soria et al. (2010), Calamita et al. (2005)

rAQP9a Amiry-Moghaddam et al. (2005)

Chloroplast NtAQP1 Uehlein et al. (2008)

Small vesicle-like

structures

TIPs and PIPs Siefritz et al. (2001), Vera-Estrella et al. (2004),

Boursiac et al. (2005)

Peribacteroid membrane GmPIP1;1 Fortin et al. (1987), Guenther et al. (2003),

MtNIP1 Catalano et al. (2004)

LjTIP1;1 Wienkoop and Saalbach (2003)aMammalian isoforms

Plant Aquaporins 7

Page 6: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Table 2 Substrates of plant MIP isoforms

Transported solute MIP isoforma References

Water Most PIP2s possess a high water permeability

SoPIP2;1 T€ornroth-Horsefield et al.

(2006)

ZmPIP2;1, ZmPIP2;5, ZmPIP2;4 Fetter et al. (2004)

In general, TIPs possess a high water permeability

ZmTIP1;1 Chaumont et al. (1998)

AtTIP1;1 Maurel et al. (1993)

NIPs possess a rather weak water permeability

OsNIP2;1 Mitani et al. (2008)

GmNIP1;1 Wallace and Roberts (2005)

AtSIP1;1, AtSIP1;2 Ishikawa et al. (2005)

Glycerol NtAQP1 Biela et al. (1999)

SsAQP1 Moshelion et al. (2002)

NtTIPa Gerbeau et al. (1999)

OsTIP1;2, OsTIP3;2, OsTIP4;1 Li et al. (2008)

PpGIP1;1 Gustavsson et al. (2005)

AtNIP1;1, AtNIP1;2 Weig and Jakob (2000)

GmNIP1;1 Dean et al. (1999), Rivers

et al. (1997)

Urea NtTIPa Gerbeau et al. (1999)

AtTIP1;1, AtTIP1;2, AtTIP2;1, AtTIP4;1 Liu et al. (2003)

CpNIP1 Klebl et al. (2003)

Lactic acid AtNIP2;1 Choi and Roberts (2007)

Antimonite AtNIP5;1, AtNIP6;1, AtNIP7;1, LjNIP5;1,

LjNIP6;1, OsNIP2;1

Bienert et al. (2008b)

AtNIP1;1 Kamiya and Fujiwara

(2009)

Arsenite OsNIP1;1,

OsNIP2;1, OsNIP2;2

Ma et al. (2008a)

AtNIP5;1, AtNIP6;1, AtNIP7;1, LjNIP5;1,

LjNIP6;1, OsNIP2;1

Bienert et al. (2008b)

AtNIP1;1 Kamiya et al. (2009)

Methylated arsenic

species

OsNIP2;1 Li et al. (2009)

Silicic acid OsNIP2;1 Ma et al. (2006)

OsNIP2;2 Yamaji et al. (2008)

ZmNIP2;1, ZmNIP2;2 Mitani et al. (2009)

HvNIP2;1 Chiba et al. (2009)Boric acid AtNIP5;1 Takano et al. (2006)

AtNIP6;1 Tanaka et al. (2008)

Hydrogen peroxide AtTIP1;1, AtTIP1;2 Bienert et al. (2007)

AtPIP2;1, AtPIP2;4 Dynowski et al. (2008a, b)

Carbon dioxide NtAQP1 Uehlein et al. (2003, 2008)

HvPIP2;1 Hanba et al. (2004)

Ammonia TaTIP2;1, TaTIP2;2 Jahn et al. (2004)

AtTIP2;1, AtTIP2;3 Loque et al. (2005)

Nitric oxide hAQP1b Herrera et al. (2006)aThis list is not exhaustivebMammalian isoform (h human)

8 G.P. Bienert and F. Chaumont

Page 7: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

2.4 Nodulin26-Like Intrinsic Proteins

One of the first identified and biochemically investigated plant MIPs was the

soybean GmNIP1;1 (Fortin et al. 1987); however, it was not recognized as a

water channel. The NIP subfamily encompasses highly divergent isoforms. This

diversity can be seen both at the amino acid sequence level and in terms of their

substrate specificities. NIPs display the greatest diversity in terms of the amino acid

residues comprising the selective filters (see below) (Bansal and Sankararamak-

rishnan 2007). This sequence diversity represents the basis for their involvement in

the transport of diverse solutes, such as water, ammonia, urea, glycerol, organic

acids, and metalloids such as boric acid, antimonite, arsenite, and silicic acid

Table 3 Biotic and abiotic factors influencing the expression of some plant MIPs

Factors influencing

expression

MIP isoform References

Water supply and osmotic

conditions

AtMIP Boursiac et al. (2005), Alexandersson et al.

(2005)

OeTIP1;1, OePIP1;1,

OePIP2;1

Secchi et al. (2007)

NtPIP1;1, NtPIP2;1 Mahdieh et al. (2008)

Hormones

Gibberilic acid PsTIP1 Ozga et al. (2002), Kolla et al. (2004)

Abscisic acid OsPIPs Lian et al. (2006)

OsTIPs Li et al. (2008)

AtPIP1 Kaldenhoff et al. (1993)

Auxin PgTIP Lin et al. (2007

Ethylene RhPIP2;1 Ma et al. (2008b)

VvPIPs Chervin et al. (2008)

Light AtPIP1 Kaldenhoff et al. (1993)

DcPIPs, DcTIPs Sato-Nara et al. (2004)

Circadian rhythm ZmPIP2s Lopez et al. (2003)

NtPIPs Siefritz et al. (2004)

Temperature ZmPIP1s, ZmPIP2s Aroca et al. (2006)

RcPIP2;1, RcPIP2;2 Peng et al. (2007)

OsPIPs Yu et al. (2006)

Nutrients AtNIP5;1 Takano et al. (2006)

OsNIP2;1 Ma et al. (2006)

Tissue-specific and

developmental

expression

ZmPIPs Hachez et al. (2006a, b, 2008), Heinen et al.

(2009), Chaumont et al. (2001)

OsPIPs Sakurai et al. (2008)

AtMIPs Alexandersson et al. (2005)

PtNIP1 Ciavatta et al. (2001)

OsNIP2;1 Ma et al. (2006)

Symbiotic and pathogen

interactions

MtNIP1 Uehlein et al. (2007)

GmNIP;1 Fortin et al. (1987)

LjTIP1 Wienkoop and Saalbach (2003)

GhMIPs Dowd et al. (2004)

PtPIPs Marjanovic et al. (2005)

Plant Aquaporins 9

Page 8: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

(Table 2). Interestingly, Arabidopsis and rice lines in which the expression of NIPgenes has been knocked out or downregulated show clear phenotypes demonstrat-

ing physiologically crucial roles of the encoded proteins in the uptake and translo-

cation of the nutritionally important metalloids boron and silicon and in the

extrusion of the highly toxic metalloids arsenite and antimonite (Takano et al.

2006; Ma et al. 2006; Yamaji et al. 2008; Isayenkov and Maathuis 2008). NIPs

are found in the endoplasmic reticulum or the plasma membrane (Table 1). NIP

expression is not as ubiquitous as that of PIPs or TIPs, but restricted to defined cell

types or tissues, consistent with their specific roles (Takano et al. 2006; Ma et al.

2006; Ciavatta et al. 2001).

2.5 X Intrinsic Proteins

As indicated by the subfamily name, XIPs are an as yet uncharacterized monophy-

letic group of MIPs from plants and fungi (Gupta and Sankararamakrishnan 2009;

Danielson and Johanson 2008). Hydropathy prediction programs suggest that XIPs

are localized to the plasma membrane of plant cells (Danielson and Johanson 2008).

The selectivity filters of XIPs from moss and fungi clearly differ chemically from

those in dicot plants, suggesting that the channeled solute and therefore the physio-

logical function will be very different. Furthermore, expression analysis indicates

that poplar XIPs do not show any tissue- or cell-specific difference in transcript

abundance (Gupta and Sankararamakrishnan 2009; Danielson and Johanson 2008).

Studies on solute transport, expression, posttranslational modifications, and other

properties will have to be performed on XIPs to shed light on this novel MIP

subfamily.

2.6 Hybrid Intrinsic Proteins

Isoforms of this subfamily have been identified in the moss P. patens and in the

spikemoss Selaginella moellendorffii, but not in vascular plants (Danielson and

Johanson 2008). HIPs seem to have been lost during the evolution of vascular

plants. As indicated by the name, HIPs share similarities with both TIPs and PIPs.

These proteins have not yet been characterized.

2.7 GlpF-Like Intrinsic Proteins

GIPs are closely related to aquaglyceroporins from bacteria (Gustavsson et al.

2005; Johanson and Danielson 2008). Two GIP sequences are known from

P. patens and another closely related moss (Gustavsson et al. 2005). One possible

explanation for the occurrence of this GLP homolog in moss species is horizontal

10 G.P. Bienert and F. Chaumont

Page 9: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

gene transfer from bacteria. Consistent with the phylogenetic relationship with

aquaglyceroporins, PpGIP1;1 heterologously expressed in Xenopus oocytes was

demonstrated to conduct glycerol, whereas only very low water permeability was

detected (Gustavsson et al. 2005).

3 Structural Features of Major Intrinsic Proteins

3.1 Aquaporin Structure

Structural information about MIPs was provided by the resolved atomic structures

of mammalian AQP1 (Murata et al. 2000; Sui et al. 2001), AQP0 (Harries et al.

2004), AQP2 (Schenck et al. 2005), AQP4 (Hiroaki et al. 2006), AQP5 (Horsefield

et al. 2008), and AQP9 (Viadiu et al. 2007), bacterial GlpF (Fu et al. 2000;

Tajkhorshid et al. 2002), AqpZ (Savage et al. 2003), and AqpM (Lee et al. 2005),

yeast PpAqy1 (Fischer et al. 2009), and protozoan PfAQP (Newby et al. 2008). To

date, spinach SoPIP2;1 is the only plant AQP for which the structure has been

resolved at atomic resolution (T€ornroth-Horsefield et al. 2006). Despite the some-

times large sequence difference between the various crystallized MIPs, the overall

structure is highly conserved (reviewed in Gonen and Walz 2006; Fu and Lu 2007).

A closer examination of the sequences of MIPs shows a tandem repeat of three

transmembrane helices and one half membrane spanning helix containing the

conserved MIP Asn-Pro-Ala (NPA) signature. The two parts of the protein are

orientated in opposite directions in the plasma membrane and form a six transmem-

brane (TM1–TM6) and two half membrane spanning helices. TM1–TM6 are

connected by five loops (loop A–loop E), the cytoplasmic loop B and the extracel-

lular loop E, forming the two short hydrophobic half-transmembrane helices. Both

termini are located in the cytoplasm. The center of this protein forms a hydrophilic

pore through the lipid bilayer, which is narrow enough to exclude the passage of

hydrated ions. The passage of nonhydrated ions is energetically unfavorable, as the

amino acids lining the pore do not allow hydrogen binding throughout the channel.

However, the carbonyl backbone of the amino acids provides hydrogen binding

partners for water or chemically similar solutes. Loops B and E, containing the

NPAmotifs, meet at the center of the pore and constitute a size exclusion zone, with

a diameter of only 3 A. The two asparagine residues constitute hydrogen bond

donors for the hydrogen acceptors of the transported molecules and their orientation

is structurally fixed in the pore. This arrangement forces the channeled compound

into an orientation, which does not allow hydrogen bond formation between

neighboring substrate molecules at the center of the channel. Substrates are there-

fore channeled in single file and possess a polarity and an electrostatic resistance,

which prevents a possible proton wire and consequently proton transport through

the pore. The available MIP structures have helped in understanding the mechanism

of transport and explaining some of the experimentally observed posttranslational

Plant Aquaporins 11

Page 10: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

modifications of MIPs from plants and mammals, such as phosphorylation and

protonation (Table 4) (T€ornroth-Horsefield et al. 2006).

3.2 Aromatic/Arginine (ar/R) Constriction Region

A second and narrower constriction region for uncharged molecules, the so-called

aromatic/arginine (ar/R) constriction region, is formed by four amino acids that

contribute to a size exclusion barrier and the hydrogen bond environment necessary

for effective transport of a substrate. One of the amino acids forming this tetrad is

in transmembrane helix 2, another in transmembrane helix 5, and the other two in

loop E. In plants, the ar/R selectivity filter is formed by a much larger number of

amino acid combinations than in mammals or microbes (Wallace and Roberts

2004; Bansal and Sankararamakrishnan 2007). Mutational studies in homologous

and heterologous expression systems have identified chemical prerequisites for the

passage of water and other compounds or the exclusion of protons (Wallace and

Roberts 2004; Beitz et al. 2006; Chen et al. 2006). In general, it can be concluded

that typical water-conducting MIPs possess rather large and hydrophilic pore

residues, which constrict the pore size to, for example, 1.86 A in AQP1. The ar/R

selectivity filter of aquaglyceroporins and NIPs is constituted of rather small and

hydrophobic residues, resulting in a larger channel diameter of 3.14 A in E. coliGlpF (EcGlpF) or even larger in NIPs.

3.3 Oligomer Formation

The structure described above is that of a MIP monomer. In membranes, MIPs

assemble as tetramers with four functional pores. In mammals, only homo-tetra-

mers are known, with the exception of hetero-tetramer formation by translational

variants using different translation start points (Tajima et al. 2010). Human AQP1 is

one of the best-studied water channels in terms of oligomerization. Three amino

acids have been shown to be essential for its tetramerization (Buck et al. 2007). In

EcGlpF, only one amino acid residue, present in the N terminus, has been demon-

strated to be critical for the proper oligomerization and in vivo stability of the

protein (Cymer and Schneider 2010).

Interestingly, data from plant PIPs indicate that several isoforms are able to form

hetero-oligomers. Maize PIP1 isoforms are inactive when heterologously expressed

in Xenopus oocytes, whereas PIP2s exhibit high water channel activity (Chaumont

et al. 2000; Fetter et al. 2004). After coexpression, ZmPIP1 and ZmPIP2 induce an

increase in the water permeability coefficient (Pf) of the membrane in a synergistic

manner compared to ZmPIP2 expression alone. When transiently and singly

expressed in maize protoplasts, ZmPIP1s and ZmPIP2s also differ in their subcel-

lular localization (Zelazny et al. 2007). ZmPIP1s are retained in the secretory

12 G.P. Bienert and F. Chaumont

Page 11: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

pathway and, more specifically, in the endoplasmic reticulum, whereas ZmPIP2s

are targeted to the plasma membrane. Upon coexpression, ZmPIP1s are localized

in the plasma membrane as a result of their physical interaction with ZmPIP2s,

as shown by F€orster resonance energy transfer analysis. Using oocytes and

Table 4 Posttranslational modifications of some plant MIP isoforms

Posttranslational

modification

MIP isoform Function References

Glycosylation McTIP1;2 Trafficking Vera-Estrella et al. (2004)

AQP2a Trafficking Hendriks et al. (2004)

Methylation Several AtPIPs Unknown Santoni et al. (2006)

N-Acetylation PvTIP3;1 Unknown Daniels and Yeager (2005)

Phosphorylation SoPIP2;1 Activity Johansson et al. (1998),

T€ornroth-Horsefield et al.

(2006)

GmNIP1;1 Activity Guenther et al. (2003)

AQP2a Trafficking Moeller et al. (2009)

ZmPIPs Activity Van Wilder et al. (2008)

þ indirect physiological data

Protonation AtPIPs Activity Tournaire-Roux et al. (2003)

AQP6a Activity Yasui et al. (1999)

AQP3a Activity Zeuthen and Klaerke (1999)

þ indirect physiological data

Ions

Zinc bAQP0a Activity Nemeth-Cahalan et al. (2007)

Lead AQP4a Activity Gunnarson et al. (2005)

Copper hAQP3a Activity Zelenina et al. (2003)

Nickel bAQP0a Activity Nemeth-Cahalan et al. (2007)

hAQP3a Activity Zelenina et al. (2003)

Mercury AtTIP1;1/AtTIP2;1 Activity Daniels et al. (1994)

hAQP1a Activity Preston et al. (1993)

Savage and Stroud (2007)

þ indirect physiological data

Silver/Gold ND Activity Niemietz and Tyerman (2002)

Calcium SoPIP2;1

PIPs in general?

Gating T€ornroth-Horsefield et al.

(2006)

þ indirect physiological data

Redox

modifications

þ indirect

physiological data

Activity Henzler et al. (2004), Ye and

Steudle (2006),

Ampilogova et al. (2006)

Heteromer

formation

PIP1 and PIP2

isoforms of several

species

Trafficking Fetter et al. (2004), Temmei

et al. (2005), Zelazny et al.

(2007), Vandeleur et al.

(2009)

Cleavage of the

N or C

terminus

AQP0a Switch from a

transport protein

to a cell

adhesion protein

Gonen et al. (2004), Scheuring

et al. (2007)

ND not determinedaMammalian isoforms (h human, b bovine)

Plant Aquaporins 13

Page 12: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

mammalian COS cells, similar results have been obtained for Mimosa pudicaMpPIP1 and MpPIP2 and Nicotiana tabacum NtPIP1;1 and NtPIP2;1 (Temmei

et al. 2005; Vandeleur et al. 2009). This hetero-oligomer formation of different PIP

isoforms has so far been observed in plant AQPs, but not mammalian or microbial

AQPs. Together, all these data suggest that the trafficking of plant AQPs plays a key

role in the currently poorly understood physiological regulation of AQP functions.

The structural basis of the physical interaction between PIPs has yet to be discovered.

The crystallization of members of the TIP, SIP, NIP, GIP, HIP, and XIP

subfamilies will probably reveal unexpected and exciting details of plant AQP

function, such as features responsible for gating and substrate selectivity, possible

motifs for posttranslational modifications, and structural cues for the formation of

oligomers.

4 Why Do Plants Contain So Many MIP Isoforms?

The previous section provided a broad picture of the multiplicity of known MIPs,

their structural features, and cellular localization. In addition, the level of expres-

sion of plant MIPs is highly regulated during development and by environmental

cues (Table 3), and the proteins are subjected to different posttranslational mod-

ifications regulating their trafficking and activity (Table 4) (reviewed in Chaumont

et al. 2005; Maurel et al. 2008; Heinen et al. 2009). However, the physiological

roles of MIPs in plants and how the discovered regulatory mechanisms are inter-

acting are still largely unknown. Much of the research effort has been devoted to the

characterization of MIPs with respect to cell water homeostasis. The roles of

several isoforms in transmembrane water fluxes at the cellular level, long-distance

translocation, and root or leaf hydraulic conductivity have been clearly demon-

strated (reviewed in Maurel et al. 2001, 2008; Kjellbom et al. 1999; Hachez et al.

2006a, b; Heinen et al. 2009). However, recent findings about the ability of plant

MIPs to channel a wide variety of small solutes suggest that these proteins are

involved not only in cellular water relations, but also in detoxification processes,

plant nutrition, and signal transduction.

This involvement of MIPs in many physiological key transport processes beside

water transport is apparent in the moss P. patens. The direct environment of this

moss, damp soil, is always supplied with water, so every cell is able to acquire

nutrients directly from the surrounding water phase and no uptake organs have been

formed during evolution. As mentioned previously, this plant expresses 23 different

MIPs clustered in the seven identified plant MIP subfamilies, PIPs, TIPs, NIPs,

SIPs, GIPs, HIPs, and XIPs (Danielson and Johanson 2008). This observation raises

the question of why an ancient organism in which every single cell is directly

connected to the surrounding water phase possesses 23 different water and glycerol

channels. One obvious answer, apart from the very finely tuned control of the

transport of a single substrate, is that MIP isoforms serve to channel different

plant metabolites.

14 G.P. Bienert and F. Chaumont

Page 13: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

In the following sections, we will describe the known substrates of MIPs that have

been identified either in transport assays in planta or in heterologous expression

studies. These substrates belong to different important substance classes, including

the vitally essential water molecule, diverse intermediate catabolic products, crucial

nutrients, and signal messenger molecules. Current views on proven and potential

functions of diverse MIPs in physiological transport processes are outlined.

4.1 MIP Function Related to Water Transport

The majority of plant MIPs are permeable to water. AQPs are high capacity

channels with the ability to channel up to 109 water molecules per channel per

second (Fujiyoshi et al. 2002). This allows a plant to tightly control water homeo-

stasis by lowering by several orders of magnitude the activation energy needed to

promote water transport across membranes. AQPs render liposome membranes and

plant membranes 10–100 times more permeable to water than membranes lacking

such channels (Dean et al. 1999; Sutka et al. 2005). AQPs have been shown to be

responsible for the high Pf value differences observed in several cell types or in

cells at different developmental stages of the same plant species (Chaumont et al.

2005). Independent of the level of the Pf value itself, the Pf is reduced after addition

of the AQP blocker HgCl2. Pf values correlate with the amount of AQPs in

heterologous expression systems or in transgenic plants (reviewed in Hachez

et al. 2006a, b). Together, these data show that AQPs regulate plant plasma

membrane permeability. In general, vacuolar membranes also have very high Pf

values (>200–500 mm/s). In plants, PIPs and TIPs represent the isoforms with the

highest water permeability and are probably “the” key players in water homeosta-

sis. The essential requirement for AQPs in water homeostasis is unquestioned.

However, specific PIP and TIP knock-out Arabidopsis lines show no obvious

developmental phenotypes (Javot et al. 2003; Sch€ussler et al. 2008; Beebo et al.

2009). This can be explained by the large number of isoforms that can probably

compensate for the loss of a single isoform. The involvement of certain isoforms in

water uptake from the soil, transcellular water transport, root-to-shoot water trans-

port, cellular water homeostasis via cytoplasmic volume buffering, maintenance of

cellular turgor, osmotic-driven growth control, and osmotic-driven organ movement

under diverse environmental conditions was deduced from AQP expression patterns

and characterization of plants overexpressing or silenced for specific isoforms

(reviewed in Hachez et al. 2006a, b; Maurel et al. 2008; Heinen et al. 2009).

It has also been hypothesized that MIPs can act as sensors for gradients of

osmotic or turgor pressure and that they transmit such information, in association

with control systems, to signaling pathways. This function is secondarily related to

the single-file water transport capacity itself and the flexible structure of MIP

tetramers. This hypothesis may be termed the “Hill hypothesis,” as it was formu-

lated by Hill et al. (2004), who proposed that the ability to sense osmotic or turgor

pressure gradients is inherent in the structure of each monomer in the tetramer.

Plant Aquaporins 15

Page 14: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

As described above, each monomer forms a narrow hydrophilic channel, which

separates the inner and outer atrium of the protein. A pressure or an osmotic

gradient across the membrane will induce a pressure gradient between the two

atria that results in water flux through the pore. Exclusion of solutes from the atrium

will create a negative physical pressure within the atrium, producing an asymmetric

deformation of the monomer or even the tetramer. Such changes in water or solute

permeability due to changes in the osmotic or hydraulic pressure have been

observed in different experimental systems and plant cell types (Wan et al. 2004;

Ye et al. 2004; Kim and Steudle 2007). These studies showed that MIPs can sense

an osmotic or pressure gradient across the membrane, probably by an asymmetric

protein conformational change, and that they are able to act as sensors of such

signals. Hill et al. (2004) suggested that MIPs are also involved in the generation of

the response by directly interacting with downstream signaling components through

a conformational change after signal perception. However, no study has yet identi-

fied such a sensor for osmotic or pressure gradients in plants, even though a general

need for such a sensor has been formulated many times and downstream signaling

pathways have been demonstrated in various organisms. Support for this MIP-

mediated gradient sensing has come from two elegant studies on the osmotic effects

on vacuolar ion release in guard cells of Commelina communis plants by MacRo-

bbie (2006a, b), who demonstrated that the tonoplast of guard cells can sense an

osmotic gradient and respond to water influx into the vacuole via increased vacuolar

ion efflux. This suggests that MIPs, especially TIPs, which are expressed in the

tonoplast of guard cells (Sarda et al. 1997), represent the most likely candidates for

the described osmotic sensor. It has been speculated that the conformational change

of TIPs caused by the pressure gradient across the channel induces ion efflux

(MacRobbie 2006b). Two possibilities were proposed to explain the ion efflux:

either TIPs become permeable to ions as a result of their deformation or TIPs are

associated in the tonoplast with ion channels that are activated in response to the

conformational change. In general, MIPs are not ion-permeable, although NOD26,

AQP1, and AQP6 have been reported to transport ions (Anthony et al. 2000; Holm

et al. 2005; Yasui et al. 1999). AQP6 shows ion permeability in intracellular vesicles

after Hg treatment (Yasui et al. 1999). This is particularly interesting, as Hg

treatment also induces ion permeability of the tonoplast in the guard cell (MacRo-

bbie 2006b). These results can be explained through a conformation change induced

by either a pressure gradient or Hg, which renders the TIP ion-permeable or induces

a signal pathway resulting in ion flux. Hill et al. (2004) and MacRobbie (2006b)

suggested that a TIP is involved in the sensing of osmotic signals; however, the

downstream signaling pathway is not known.

4.2 MIP Function Related to Nitric Oxide Transport

It was recently discovered that MIPs facilitate the transmembrane transport of nitric

oxide (NO), a hydrophobic gas that plays an important signaling role in a variety of

16 G.P. Bienert and F. Chaumont

Page 15: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

physiological process in all organisms. Using an NO-specific fluorescent dye,

Herrera et al. (2006) directly correlated the rate of NO influx into mammalian

endothelial cells and liposomes with the abundance of AQP1. In plants, NO triggers

signaling pathways involved in programmed cell death, pathogen defense, flower-

ing, stomatal closure, and gravitropism (Neill et al. 2008). As earlier studies had

reported that mammalian membranes represent no significant barrier for its diffu-

sion, NO was thought to move from its site of production to its site of action by

freely diffusing through the lipid bilayer of cell membranes without a need for a

specific transporter (Liu et al. 2002). A study using various simulation techniques to

examine the NO permeability of brain AQP4 suggested that the central pore formed

by the association of the four monomers in a tetramer presents a lower barrier to NO

gas permeation than pure lipid bilayers (Wang and Tajkhorshid 2010). The authors

suggested that the central pore of AQP4 tetramer may act as a reservoir for NO

messenger molecules, but doubted that AQP4 plays a physiological role in trans-

membrane NO transport, except in membranes with a low intrinsic gas permeability

or when AQPs occupy a large portion of the membrane surface. The questions of

whether plant MIPs are involved in the controlled distribution of the signaling

molecule NO and whether plant membranes present a higher resistance to NO

diffusion than mammalian membranes remain to be answered. Plant PIPs resemble

the most to AQP1-like proteins from their structure and transport specificity. It will

be interesting to investigate the permeability and the possible role of PIPs in NO

signaling in MIP-deficient mutant plants and in transport assays. Future studies will

show whether MIPs represent a key signal pore for NO, or whether this capacity

represents a peculiar, but irrelevant, side effect.

4.3 MIP Function Related to Ammonia Transport

Approximately 1% of all man-made energy is used to produce ammonia gas, the

direct or indirect precursor of most nitrogen-containing compounds. At physiologi-

cal pH, the uncharged ammonia molecule is in thermodynamic equilibrium with its

corresponding acid, the ammonium ion (NH4þ). All assimilated nitrogen in plant

cells is transformed into NH4þ, the precursor for amino acids and all other

N-containing molecules. NH4þ is transported across plant membranes via the

ammonium protein transporter family. NH3 has been suggested to cross membranes

by free diffusion. However, the protein-mediated transport of NH3 has some

advantages compared to that of NH4þ. Potassium ions, which interfere with the

binding sites of the chemically similar NH4þ ion in transporter proteins, do not

compete with NH3 for transport. Additionally, transport can occur without the use

of energy, as it is electrically neutral.

A yeast strain lacking its three intrinsic ammonium transporters and thus

unable to grow on medium with low nitrogen levels was transformed with a

cDNA library from wheat with the aim of identifying new transporters for NH4þ

or NH3 (Jahn et al. 2004). In addition to some already known high affinity

Plant Aquaporins 17

Page 16: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

ammonium transporters, low affinity transporters belonging to the TIP subfamily

were identified in this screen. In a similar approach, TIPs from Arabidopsis wereshown to be permeable to NH3 (Loque et al. 2005). NH3 has comparable dimen-

sions to water and its dipole moment is very close to that of water, suggesting that

both may use the same channel (Jahn et al. 2004). A variety of MIPs were then

shown to facilitate the transport of NH3 in various heterologous test systems, and

structural prerequisites for the ar/R selectivity filter were identified (Beitz et al.

2006; Dynowski et al. 2008a, b). In plants, nitrogen is one of the most limiting

factors for growth. NH3 gas is probably responsible for nitrogen loss from plants

(Husted and Schjoerring 1996), so mechanisms preventing this are highly benefi-

cial. Vacuoles could represent a major storage organelle where TIPs facilitate the

uptake of NH3, which is consequently trapped by being converted to the NH4þ

anion in the acidic vacuolar environment. At low cytoplasmic NH4þ/NH3 concen-

trations, the vacuole might provide a nitrogen source for metabolism in the form of

NH4þ/NH3. However, direct physiological data from TIP knockout mutants reveal-

ing such involvement of TIP-mediated NH3 membrane transport are lacking

(Sch€ussler et al. 2008). One reason for this lack of phenotype could be due to

compensation mechanisms of closely related isoforms. The ability of several TIPs

to channel NH3 seems to be conserved across plant species and probably plays an

important role in nitrogen transport.

4.4 MIP Function Related to Urea Transport

Urea is a secondary nitrogen metabolite in plants and has the highest nitrogen

content of commonly used solid nitrogenous fertilizers. This lowest transportation

cost per unit of nitrogen nutrient, together with its high solubility in water and

readily assimilation by plants, makes urea a useful agronomical fertilizer. In

Arabidopsis, a single gene codes for an active urea transport system, the proton–

urea cotransporter AtDUR3 (Kojima et al. 2006). When urea is the sole nitrogen

source in the medium, the AtDUR3 knockout line becomes chlorotic compared to

the wild type, in which AtDUR3 transcript levels increase markedly under nitro-

gen limiting conditions (Kojima et al. 2007). How urea is relocalized and trans-

located across membranes for use as a nitrogen source is mostly unknown. As

urea is uncharged, it diffuses slowly across membranes, and low affinity channels

could be important for facilitated relocation between different compartments.

Members of the NIP, PIP, and TIP subfamilies have been shown to facilitate

the transport of urea across membranes (Gerbeau et al. 1999; Klebl et al. 2003;

Liu et al. 2003; Wallace and Roberts 2005). Vacuoles might be used for short-

term storage of urea or to avoid toxic concentrations in the cytoplasm. Phenotypes

associated with TIP-mediated urea transport across the tonoplast of Arabidopsishave not been reported.

18 G.P. Bienert and F. Chaumont

Page 17: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

4.5 MIP Function Related to Carbon Dioxide Transport

Life on earth became possible through the evolution of photosynthesis and plant

carbon fixation. Rubisco, a chloroplast bifunctional enzyme that catalyzes the

transfer of CO2 to ribulose 1,5 biphosphate in the Calvin cycle, requires a certain

CO2 threshold concentration to be efficient, and the CO2 concentration in the

vicinity of the enzyme is therefore the limiting factor for photosynthesis. It is

assumed that the average concentration of CO2 in chloroplasts is only half of the

atmospheric CO2 concentration of 0.038%. The diffusion of CO2 from the atmo-

sphere into the substomatal cavity via the stomata and to sites of carboxylation via

the mesophyll is the main factor controlling CO2 availability for Rubisco. Meso-

phyll conductance has a greater impact than stomatal conductance on the final

concentration in the chloroplast and is regulated by the diffusion barriers for CO2

formed by the intercellular space, apoplastic liquid phase, plasma membrane,

cytosol, chloroplast membranes, and stroma. Plants are able to facilitate the diffu-

sion of CO2 into the chloroplast. However, the mesophyll conductance can be

rapidly altered by several environmental factors, such as light, temperature,

water, and CO2 (Flexas et al. 2008). Another way to influence mesophyll conduc-

tance would be a change in the CO2 permeability of the plasma membrane and

chloroplast membranes. Recently, much attention has been paid to the involvement

of MIPs in CO2 conduction across cellular membranes. The first evidence for this

came from a study by Terashima and Ono (2002), who showed that the AQP

inhibitor HgCl2 reduces the mesophyll conductance for CO2. Thereafter, it was

shown in the heterologous oocyte system that NtAQP1, a PIP1 isoform from

tobacco, is able to significantly facilitate the transmembrane diffusion of CO2

(Uehlein et al. 2003). Tobacco plants overexpressing NtAQP1 show increased net

photosynthesis and leaf growth, whereas NtAQP1-silenced plants show decreased

net photosynthesis and mesophyll conductance (Uehlein et al. 2003, 2008). How-

ever, the change in conductivity may not be explained solely by the modulated CO2

membrane permeability due to AQPs, as a change in AQP quantity and localization

could also influence water homeostasis by changing stomatal conductance. Uehlein

et al. (2008) showed that, in tobacco, NtAQP1 is localized to the inner chloroplast

envelope in addition to the plasma membrane. Water conductivity tests performed

on isolated chloroplast or plasma membrane vesicles showed that the plasma

membrane water permeability of NtAQP1-silenced plants is reduced to approxi-

mately 50% of that in wild type tobacco plants, while chloroplast water permea-

bility is not changed. In contrast, the chloroplast envelope CO2 permeability of

NtAQP1-silenced plants is markedly decreased compared to that in the wild type,

while the plasma membrane CO2 permeability is unaffected. Additionally, this

study indicated that the plasma membrane is about five times more permeable to

CO2 than the chloroplast envelope. It was concluded that the chloroplast envelope

represents a higher diffusion barrier than previously estimated and that NtAQP1

markedly affects chloroplast CO2 permeability in tobacco (Uehlein et al. 2008). An

increased CO2 conductance in plants induced by PIP overexpression has also been

Plant Aquaporins 19

Page 18: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

reported in rice (Hanba et al. 2004). Physiologically relevant MIP-mediated CO2

transport across mammalian membranes is more controversial. While there is no

debate about the fact that CO2 can pass through mammalian AQPs, as shown in

several transport and simulation studies, the question is whether AQPs can increase

the diffusion of CO2 across lipid membranes which, in the so far investigated

mammalian systems, have an already very high intrinsic CO2 permeability (Fang

et al. 2002). In addition, the measured resistance to transmembrane CO2 permeation

in mammalian cell systems has been reported to be mainly caused by an unstirred

layer next to the membrane and not by the expressed protein (e.g., AQPs) levels

therein or the membrane composition itself (Missner et al. 2008). Physiological and

molecular data for PIP-mediated CO2 transport across the plasma membrane and

chloroplast envelope point more toward a potential role of these channels in CO2

transport in plants. Nevertheless, it remains to be determined whether the composi-

tion of the plant membrane causes a higher intrinsic CO2 resistance than that seen

with mammalian membranes.

4.6 MIP Function Related to Hydrogen Peroxide Transport

On the one hand, various studies have demonstrated the essential function of H2O2 as

a signaling molecule controlling an amazingly wide spectrum of processes in plants

(Neill et al. 2002). On the other hand, H2O2 is a reactive oxygen species (ROS), an

oxidant that can react with various cellular targets and cause cell damage or even cell

death. Cellular levels of H2O2 have therefore to be tightly regulated. The average

half-life of H2O2 is long enough for it to act as a transportable signal, but short

enough for its concentration to fluctuate and represent a potent on–off modulator. Its

concentration in plant tissues is in the micromolar to low millimolar range, and its

stability, in combination with its chemical reactivity, makes H2O2 a good multi-

facetted messenger molecule. From the studies on yeast and bacteria, it is clear that

the diffusion of H2O2 across membranes is limited, as they are quite impermeable to

H2O2 (reviewed in Bienert et al. 2006). This permeability needs to be adjustable to

allow signaling. It has been shown that yeast alters its permeability to H2O2 by a

change in membrane composition (Branco et al. 2004; Matias et al. 2007). Like yeast

or bacteria, mammals and plants have H2O2-impermeable membranes (Fritz et al.

2007). The transmembrane movement of H2O2 through AQPs was proposed in 2000

as a result of a biophysical study (Henzler and Steudle 2000). A broad survey aiming

at the molecular identification of H2O2-permeable MIPs revealed that plant TIP1 and

mammalian AQP8 are highly permeable to H2O2. Later, the ability to transport H2O2

was also demonstrated for a variety of other MIP isoforms (AtTIP2;3, AtNIP1;2,

AtPIP2;1, and AtPIP2;4) from different plant subfamilies after expression in yeast

(Dynowski 2008a, b). Molecular dynamic simulation analysis confirmed the H2O2

permeability of MIPs. H2O2 shares several chemical features with water (size, elec-

trochemical properties, and ability to form hydrogen bonds), which makes them both

candidates of MIP substrates (Bienert et al. 2006). Because of the different pore

20 G.P. Bienert and F. Chaumont

Page 19: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

layouts of different MIPs, the level of H2O2 permeability varies considerably. While

some isoforms, such as the TIP1s, are highly permeable to H2O2, TIP2s and PIPs are

less permeable. A mutational approach revealed that amino acid mutations in

AtPIP2;1 that abolish water permeability also abolish H2O2 permeability, suggesting

that water and H2O2 share the same pathway through AQPs. Although the ability of

specific MIPs to mediate H2O2 transport has been analyzed in detail, the physiologi-

cal relevance of the MIP-mediated transport of this signaling molecule in plants and

mammals is not yet resolved. Analysis of catalase activity and anthocyanin content in

an Arabidopsis double mutant in which the H2O2-permeable TIP1;1 and TIP1;2

isoforms were both knocked out revealed a minor change in the steady-state level

compared to the wild type under nonstressed conditions (Sch€ussler et al. 2008),suggesting a potential involvement of TIPs in redox metabolism. However, as the

difference was minor, it is hard to make any conclusions from these observations.

The presence of MIPs facilitating the diffusion of H2O2 across the tonoplast might

seem surprising, given the presence of several effective cytoplasmic detoxification

systems (reviewed in Mittler 2002 and Smirnoff 2005). However, vacuolar flavo-

noids have been shown to have highly antioxidative properties in vitro and in vivo

and, in combination with vacuolar peroxidases, vacuoles potentially represent an

effective ROS detoxification system (Smirnoff 2005). In this context, some very

interesting observations were made in cytochemical studies. In pea plants, cadmium

stress-induced H2O2 has been clearly detected in the tonoplast (Romero-Puertas

et al. 2004). Similar “deposition” of H2O2 was seen in the tonoplast of leaf sheath

cells of salt-stressed rice plants (Wi et al. 2006). These observations might represent

snapshots of transmembrane H2O2 transport through TIP proteins.

In salt-stressed Arabidopsis plants, the formation of intracellular vesicles accu-

mulating H2O2 can be observed in root cells (Leshem et al. 2006, 2007). These

vesicles are subsequently targeted to the tonoplast, delivering the H2O2 into the

vacuole. Interestingly, in the same conditions, Arabidopsis relocalized TIP1;1 pro-

teins into intravacuolar vesicles (Boursiac et al. 2005). As TIP1;1 is permeable to

H2O2 this internalization could prevent the efflux of H2O2 back into the cytoplasm.

In agreement with this hypothesis, AtTIP2;1, which is less permeable to H2O2,

was not internalized (Boursiac et al. 2005). However, this mechanistic interplay of

H2O2-filled vesicles and AQP redistribution is far from being understood.

In terms of the involvement of MIPs in the regulation of plasma membrane

permeability for H2O2, it will be interesting to study mutant plants in which H2O2-

permeable PIPs are knocked out. Investigations on intercellular and membrane-

crossing H2O2 signal transduction pathways and cell detoxification mechanisms

should be prioritized.

4.7 MIP Function Related to Organic Acid Transport

Small organic acids represent one class of solutes that has been known for a long

time to permeate through mammalian MIPs and speculated to be transported by

Plant Aquaporins 21

Page 20: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

plant AQPs (Tsukaguchi et al. 1998; Ouyang et al. 1991). Rat AQP9 is permeable to

lactic acid in a pH-sensitive manner (Tsukaguchi et al. 1998). This lactic acid

permeability at slightly acidic pH might be important during lactic acidosis in brain

ischemia, in which AQP9 may contribute significantly to the clearance of excess

lactate from the extracellular space under pathological conditions (Badaut et al.

2001). Recently, the aquaglyceroporin Fps1p from Saccharomyces cerevisiae wasshown to mediate the transport of the undissociated form of acetic acid across yeast

membranes (Mollapour and Piper 2007). Toxic external concentration levels of

acetic acid trigger Hog1p-mediated phosphorylation of Fps1p, which results in the

degradation of Fps1p in the vacuole, clearly indicating the involvement of this

aquaglyceroporin in the regulation of acetic acid levels. Bacteria, fungi, and

vertebrates contain isoforms of a monocarboxylate transport protein family,

which is, interestingly, absent in plants (Halestrap and Price 1999). The expression

of AtNIP2;1 from Arabidopsis is upregulated 300-fold after anoxic stress, in which

lactic acid fermentation precedes ethanolic fermentation. When heterologously

expressed in oocytes, NIP2;1 was shown to be highly permeable to lactic acid,

but only poorly permeable to water and glycerol (Choi and Roberts 2007). It is,

therefore, tempting to speculate that AtNIP2;1 is important for the efflux of lactic

acid under anoxic conditions. The first NIP described, namely GmNOD26

(GmNIP1;1), was suggested to be permeable to malate (Ouyang et al. 1991), as

its phosphorylation status correlated closely with malate uptake across the peri-

bacteroid membrane. Recently, a peptide transporter from Alnus glutinosa was

identified as the first organic acid transporter in a peribacteroid membrane (Jeong

et al. 2004). Nevertheless, an additional low affinity channel, such as a MIP,

would result in a highly efficient uptake system. Further investigations on the

ability of MIPs to channel organic acids are required. In particular, NIPs and the

uncharacterized XIPs, which have rather wide and hydrophobic selectivity filters,

represent potential candidates for such transporters.

4.8 MIP Function Related to Glycerol Transport

Glycerol is a triglyceride molecule that constitutes building blocks of biological

membranes. In addition to its structural function, it is involved in energy metabo-

lism as glycerol trisphosphate. Aquaglyceroporins from mammals and microbes

have been demonstrated to play important roles in osmotic or freezing tolerance,

regulation of the glycerol content of diverse tissues, and fat metabolism (Tamas

et al. 1999; Rojek et al. 2008). Members from various plant MIP subfamilies, such

as PpGIP1;1 from Physcomitrella (Gustavsson et al. 2005), and several TIP and PIP

isoforms can facilitate the transport of glycerol across membranes (Gerbeau et al.

1999; Siefritz et al. 2001). Initially, NIP isoforms were characterized as predomi-

nantly glycerol facilitators and, for a long time, were seen as functional counter-

parts to aquaglyceroporins from microbes and mammals (Weig and Jakobs 2000).

However, the lack of physiological evidence for NIP-mediated glycerol transport in

22 G.P. Bienert and F. Chaumont

Page 21: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

plants and the conclusive evidence for NIP function in relation to metalloid

transport make a main role of NIPs in glycerol transport unlikely. Glycerol is a

very flexible molecule, with 576 known conformations (Law et al. 2005). It has

been speculated that the ability of NIPs to channel glycerol is a simple consequence

of the structural similarity of glycerol to metalloids, rather than to a physiological

function (Porquet and Filella 2007; Bienert et al. 2008a).

4.9 MIP Function Related to Boric Acid Transport

Boron (B) has a special position among the essential elements. First, the difference

between deficient and toxic concentrations in the soil solution is smaller than for

any other nutrient element, and, second, it is the only nutrient that is assimilated as

an uncharged molecule, boric acid. Its uptake and translocation have therefore to be

highly regulated. Physiological experiments have demonstrated a channel-mediated

boron transport system in plants. Nevertheless, the first experimental proof of a

B transporter in plants revealed an active B efflux transporter (BOR1) with a high

similarity to anion exchanger proteins (Takano et al. 2002). BOR1 homologs were

also shown to be efflux transporters translocating B within the plant (Takano et al.

2008). Recently, the channel protein for boric acid uptake was identified in Arabi-dopsis. AtNIP5;1 proved to be essential for B uptake under B-limiting conditions

(Takano et al 2006). In a microarray analysis, expression of AtNIP5;1was shown tobe induced under low B conditions. The B transporting ability of AtNIP5;1 has been

demonstrated in vivo and in heterologous expression systems. The cooperative

functional interaction between AtNIP5;1, which takes up boric acid from the

soil, and AtBOR1, which loads borate into the xylem, was shown to result in

effective boron nutrition for the plant. In accordance with these results, only

co-overexpression of AtBOR1 in an AtNIP5;1 activation tag line resulted in plants

with a super-tolerance to a low B concentration (Kato et al. 2009). In contrast,

overexpression of AtNIP5;1 under the control of the CaMV 35S promoter resulted

in reduced overall growth. These results make clear that the overexpression of

a gene without respecting the physiological context might lead to unexpected

results. In this case, only the enhanced expression of AtNIP5;1 in cells in which

it is endogenously expressed combined with modulation of AtBOR1 expression

resulted in a positive outcome.

AtNIP6;1, the most similar gene to AtNIP5;1, is also permeable to boric acid,

but possesses no water conductivity (Tanaka et al. 2008). Under conditions of B

deprivation, AtNIP6;1 expression, mainly in nodal regions, is only increased in

shoots, and AtNIP6;1 knockouts only show reduced rosette leaf expansion and

decreased B concentrations under low B conditions. This suggests that NIP6;1

plays a role in the xylem–phloem transfer of boric acid in the nodal regions. It is

worth mentioning that AtNIP5;1 and AtNIP6;1 evolved differently in terms of their

expression pattern and substrate specificity, apart from that for boric acid.

Plant Aquaporins 23

Page 22: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

4.10 MIP Function Related to Silicic Acid Transport

Surprisingly, the first silicon (Si) transporter discovered in vascular plants was

identified as a MIP, more precisely OsNIP2;1 from rice (Ma et al. 2006). Si, the

second most abundant element in the Earth’s crust, is essential for animals and

diatoms and highly beneficial in a wide range of plant species. Si increases the

tolerance of certain plants to an impressive variety of biotic and abiotic stresses and

also increases their mechanical strength. In addition to its structural support func-

tion, Si has been suggested to interact with stress-related signaling systems to

increase plant resistance to pathogens (Fautteux et al. 2005). It is taken up by the

roots in the form of the bioavailable silicic acid [Si(OH)4], a quite large uncharged

molecule with restricted free diffusion across the lipid bilayer. After a long search,

the genes encoding Si transporters were identified in a set of rice mutants defective

in Si uptake. Ma and coworkers showed that a concerted cooperation of influx and

efflux transporters is responsible for the uptake and translocation of Si in monocot

species (Ma et al. 2006, 2007). While Si influx is mediated by MIPs, the efflux is

mediated by proteins with similarities to anion efflux transporters. The roles of

different MIP isoforms from rice, maize, and barley in Si uptake and distribution

were unraveled, in part, using mutants (Ma et al. 2006). It will be interesting to

determine which proteins are responsible for Si transport in dicot plants, as homo-

logs of the Si transporting monocot NIPs have not yet been identified in dicots. The

fact that the functions of MIPs, initially identified as essential water transporters,

have been expanded to nutritionally important transport systems for large

uncharged molecules, such as Si(OH)4, points out their fundamental importance

for life, especially in those plants that have evolved multiple isoforms.

4.11 MIP Function Related to Arsenite/Antimonite Transport

Aquaglyceroporins from diverse organisms have been shown to be involved in

physiological important transmembrane glycerol transport. A comparison of glyc-

erol, arsenite [As(OH)3], and antimonite [Sb(OH)3] revealed their high similarity in

structural and electrochemical properties, which are also important for membrane

transport (Porquet and Filella 2007). It was concluded that a MIP channel protein

that is permeable for glycerol should also transport As(OH)3 and Sb(OH)3. The

subsequent discovery that aquaglyceroporins play physiological roles in resistance

to the highly toxic metalloid arsenic was therefore only a matter of time. The

legume symbiont Sinorhizobium meliloti takes up arsenate via phosphate transpor-

ters and the arsenate is reduced to As(OH)3 in the cytosol by the arsenate reductase

ArsC. As(OH)3 is transported by the bacterial MIP AqpS down the chemical

concentration gradient out of the cell. AqpS, together with all other components

involved in arsenic resistance, is encoded in the arsenic resistance operon

(Yang et al. 2005). Several protozoan parasites decrease their expression of

24 G.P. Bienert and F. Chaumont

Page 23: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

aquaglyceroporins and become resistant to arsenite- or antimonite-containing

drugs, normally used as first line treatments (Mukhopadhyay and Beitz 2010).

The increased resistance correlates with a reduced level of the metalloid in the

cells. The parasite aquaglyceroporins were demonstrated to be permeable to these

metalloids in transport assays.

The yeast aquaglyceroporin Fps1 allows the bidirectional flux of As(OH)3 across

the yeast membrane. It was shown that Fps1p mediates As(OH)3 efflux in concert

with another arsenic transporter, Acr3p, and plays a key role in the arsenic tolerance

of the yeast. Yeast cells use Fps1-mediated As(OH)3 efflux to ensure efficient

detoxification of the cytoplasm when As(OH)3 accumulates inside cells and prevent

the detrimental influx of As(OH)3 via posttranslational gating of Fps1, the closure

of Fps1 being mediated by the Hog1-dependent kinase signaling cascade (Thorsen

et al. 2006).

The variety of pore layouts in NIPs suggests that they may represent the until

recently uncharacterized arsenite channels in plants. Four studies have indepen-

dently demonstrated NIP-mediated As(OH)3 transport in heterologous systems and

directly in plants, using knockout mutants of rice and Arabidopsis (Bienert et al.2008b; Isayenkov and Maathuis 2008; Kamiya et al. 2009; Ma et al. 2008a). Yeast

growth and survival assays and direct uptake assays in both yeast and oocytes

showed that NIP isoforms from Arabidopsis, rice, and Lotus japonicus from all

three phylogenetic NIP groups are permeable to As(OH)3. AtNIP1;1 knockout

mutants of Arabidopsis show increased resistance when grown on medium contain-

ing As(OH)3 (Kamiya et al. 2009) and a field-grown lsi1 mutant of rice had lower

As concentrations in the straw than the wild type, but roughly the same levels in the

grain and husk (Ma et al. 2008a). AtNIP1;2, AtNIP5;1, AtNIP6;1, and AtNIP7;1knockout lines do not show any increased tolerance to As(OH)3, although they have

been shown to be permeable to As(OH)3 when expressed in heterologous systems

(Bienert et al. 2008; Kamiya et al. 2009). These results revealed once again that a

shared substrate specificity of proteins does not necessarily result in similar physi-

ological functions. Although total arsenic uptake might be unchanged in a mutant

knocked-out for one MIP channel compared to the wild type, translocation and

subsequently compartmentalization could differ significantly. The characterization

of more NIP genes in one plant species will provide a better understanding of the

molecular mechanisms of As(OH)3 fluxes in plants.

Given that, in contrast to the situation in prokaryotes and other eukaryotes,

several eukaryotes, including plants, do not, as far as we are aware, encode specific

active arsenite efflux transporters, but do contain arsenate reductases and metalloid-

permeable MIPs and are mainly exposed to arsenate, rather than arsenite, it is

tempting to hypothesize that MIPs constitute a major arsenic detoxification path-

way. One physiological study in rice and tomato provided evidence for such a high

capacity As(OH)3 efflux system in roots (Xu et al. 2008). Recently, Zhao et al.

(2010) showed that OsNIP2;1 accounts for 15–20% of the total As(OH)3 efflux

from rice roots, clearly indicating that plant AQPs are involved in arsenic detoxifi-

cation processes. In addition to their involvement in the transport of arsenic, several

NIP isoforms have been shown to be permeable to the toxic metalloid Sb(OH)3 in a

Plant Aquaporins 25

Page 24: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

heterologous yeast system and in planta (Bienert et al. 2008; Kamiya and Fujiwara

2009). Nevertheless, proof is required that certain NIP isoforms play a role in plant

arsenic and antimony transport metabolism and detoxification and that this ability is

not an accidental side effect due to chemical and structural similarity to other

metalloids.

5 Conclusion

Taking together all the data on the transport abilities of MIPs, it seems that, during

evolution, there was a general trend to the production and the high conservation of

two constriction regions (NPA boxes and ar/R selectivity filter), in which the pore-

lining residues generate an environment with very strict selectivity against certain

substrate groups, such as protons and other ions. In contrast, the concomitant

selectivity for, or against, uncharged solutes was rather flexible and not that

stringent. This selectivity against charged, but not uncharged, molecules, together

with the multiplication of isoforms in plant species, opened up the broad substrate

spectrum of today’s MIPs. This, together with the array of physical and biochemical

channel properties, was the decisive factor in their involvement in many physiolog-

ically important transport processes, including water homeostasis, plant nutrition,

logistics and circulation of plant metabolites, and signaling processes. However,

further investigation of the function and regulation of plant MIPs is still required.

For instance, it is particularly important to evaluate the contribution of MIPs to

signaling processes as sensors or channels transporting signaling molecules during

specific physiological processes. The deciphering of the molecular and cellular

mechanisms regulating these channels and in the proper transmission of signals is

definitely required.

Acknowledgements This work was supported by grants from the Belgian National Fund for

Scientific Research (FNRS), the Interuniversity Attraction Poles Program–Belgian Science Policy,

and the “Communaute francaise de Belgique–Actions de Recherches Concertees.” GPB was

supported by an individual Marie Curie European fellowship.

References

Agre P, Preston GM, Smith BL, Jung JS, Raina S, Moon C, Guggino WB, Nielsen S (1993)

Aquaporin CHIP: the archetypal molecular water channel. Am J Physiol 265:F463–F476

Alexandersson E, Fraysse L, Sj€ovall-Larsen S, Gustavsson S, Fellert M, Karlsson M, Johanson U,

Kjellbom P (2005) Whole gene family expression and drought stress regulation of aquaporins.

Plant Mol Biol 59:469–484

Amiry-Moghaddam M, Lindland H, Zelenin S, Roberg BA, Gundersen BB, Petersen P, Rinvik E,

Torgner IA, Ottersen OP (2005) Brain mitochondria contain aquaporin water channels:

evidence for the expression of a short AQP9 isoform in the inner mitochondrial membrane.

FASEB J 19:1459–1467

26 G.P. Bienert and F. Chaumont

Page 25: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Ampilogova YN, Zhestkova IM, Trofimova MS (2006) Redox modulation of osmotic water

permeability in plasma membranes isolatedfrom roots and shoots of pea seedlings. Russ J

Plant Physiol 53:622–628

Anthony TL, Brooks HL, Boassa D, Leonov S, Yanochko GM, Regan JW, Yool AJ (2000) Cloned

human aquaporin-1 is a cyclic GMP-gated ion channel. Mol Pharmacol 257:576–588

Aroca R, Amodeo G, Fernadez-lliescas S, Herman EM, Chaumont F, Chrispeels MJ (2005) The

role of aquaporins and membrane damage in chilling and hydrogen peroxide induced changes

in the hydraulic conductance of maize roots. Plant Physiol 137–353

Aroca R, Ferrante A, Vernieri P, Chrispeels MJ (2006) Drought, abscisic acid and transpiration

rate effects on the regulation of PIP aquaporin gene expression and abundance in Phaseolusvulgaris plants. Ann Bot (Lond) 98:1301–1310

Badaut J, Hirt L, Granziera C, Bogousslavsky J, Magistretti PJ, Regli L (2001) Astrocyte-specific

expression of aquaporin-9 in mouse brain is increased after transient focal cerebral ischemia.

J Cereb Blood Flow Metab 21:477–482

Bansal A, Sankararamakrishnan R (2007) Homology modeling of major intrinsic proteins in

rice, maize and Arabidopsis: comparative analysis of transmembrane helix association and

aromatic/arginine selectivity filters. BMC Struct Biol 7:27

Barkla BJ, Vera-Estrella R, Pantoja O, Kirch HH, Bohnert HJ (1999) Aquaporin localization –

how valid are the TIP and PIP labels? Trends Plant Sci 4:86–88

Beebo A, Thomas D, Der C, Sanchez L, Leborgne-Castel N, Marty F, Schoefs B, Bouhidel K

(2009) Life with and without AtTIP1;1, an arabidopsis aquaporin preferentially localized in the

apposing tonoplasts of adjacent vacuoles. Plant Mol Biol 70:193–2009

Beitz E, Wu B, Holm LM, Schultz JE, Zeuthen T (2006) Point mutations in the aromatic/arginine

region in aquaporin 1 allow passage of urea, glycerol, ammonia, and protons. Proc Natl Acad

Sci USA 103:269–274

Biela A, Grote K, Otto B, Hoth S, Hedrich R, Kaldenhoff R (1999) The Nicotiana tabacum plasma

membrane aquaporin NtAQP1 is mercury-insensitive and permeable for glycerol. Plant J

18:565–570

Bienert GP, Schjoerring JK, Jahn TP (2006) Membrane transport of hydrogen peroxide. Biochim

Biophys Acta 1758:994–1003

Bienert GP, Møller AL, Kristiansen KA, Schulz A, Møller IM, Schjoerring JK, Jahn TP (2007)

Specific aquaporins facilitate the diffusion of hydrogen peroxide across membranes. J Biol

Chem 282:1183–1192

Bienert GP, Sch€ussler MD, Jahn TP (2008a) Metalloids: essential, beneficial or toxic? Major

intrinsic proteins sort it out. Trends Biochem Sci 33:20–26

Bienert GP, Thorsen M, Sch€ussler MD, Nilsson HR, Wagner A, Tamas MJ, Jahn TP (2008b)

A subgroup of plant aquaporins facilitate the bi-directional diffusion of As(OH)3 and Sb(OH)3across membranes. BMC Biol 6:26

Boursiac Y, Chen S, Luu DT, Sorieul M, van den Dries N, Maurel C (2005) Early effects of

salinity on water transport in arabidopsis roots. Molecular and cellular features of aquaporin

expression. Plant Physiol 139:790–805

Branco MR, Marinho HS, Cyrne L, Antunes F (2004) Decrease of H2O2 plasma membrane

permeability during adaptation to H2O2 in Saccharomyces cerevisiae. J Biol Chem

279:6501–6506

Buck TM, Wagner J, Grund S, Skach WR (2007) A novel tripartite motif involved in aquaporin

topogenesis, monomer folding and tetramerization. Nat Struct Mol Biol 14:762–769

Calamita G, Ferri D, Gena P, Liquori GE, Cavalier A, Thomas D, Svelto M (2005) The inner

mitochondrial membrane has aquaporin-8 water channels and is highly permeable to water.

J Biol Chem 280:17149–17153

Catalano CM, Lane WS, Sherrier DJ (2004) Biochemical characterization of symbiosome mem-

brane proteins from Medicago truncatula root nodules. Electrophoresis 25:519–531

Chaumont F, Barrieu F, Herman EM, Chrispeels MJ (1998) Characterization of a maize tonoplast

aquaporin expressed in zones of cell division and elongation. Plant Physiol 117:1143–1152

Plant Aquaporins 27

Page 26: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Chaumont F, Barrieu F, Jung R, Chrispeels MJ (2000) Plasma membrane intrinsic proteins from

maize cluster in two sequence subgroups with differential aquaporin activity. Plant Physiol

122:1025–1034

Chaumont F, Barrieu F, Wojcik E, Chrispeels MJ, Jung R (2001) Aquaporins constitute a large and

highly divergent protein family in maize. Plant Physiol 125:1206–1215

Chaumont F, Moshelion M, Daniels MJ (2005) Regulation of plant aquaporin activity. Biol Cell

97:749–764

Chen H, Wu Y, Voth GA (2006) Origins of proton transport behavior from selectivity domain

mutations of the aquaporin-1 channel. Biophys J 90:73–75

Chervin C, Tira-Umphon A, Terrier N, Zouine M, Severac D, Roustan JP (2008) Stimulation of the

grape berry expansion by ethylene and effects on related gene transcripts, over the ripening

phase. Physiol Plant 134:534–546

Chiba Y, Mitani N, Yamaji N, Ma JF (2009) HvLsi1 is a silicon influx transporter in barley. Plant J

57:810–818

Choi WG, Roberts DM (2007) Arabidopsis NIP2;1, a major intrinsic protein transporter of lactic

acid induced by anoxic stress. J Biol Chem 282:24209–24218

Ciavatta VT, Morillon R, Pullman GS, Chrispeels MJ, Cairney J (2001) An aquaglyceroporin is

abundantly expressed early in the development of the suspensor and the embryo proper of

loblolly pine. Plant Physiol 127:1556–1567

Cymer F, Schneider D (2010) A single glutamate residue controls the oligomerization, function,

and stability of the aquaglyceroporin GlpF. Biochemistry 49:279–286

Daniels MJ, Yeager M (2005) Phosphorylation of aquaporin PvTIP3;1 defined by mass spectrom-

etry and molecular modeling. Biochemistry 44:14443–14454

Daniels MJ, Mirkov TE, Chrispeels MJ (1994) The plasma membrane of Arabidopsis thalianacontains a that is a homolog of the tonoplast water channel protein TIP. Plant Physiol

106:1325–1333

Danielson JA, Johanson U (2008) Unexpected complexity of the aquaporin gene family in the

moss Physcomitrella patens. BMC Plant Biol 8:45

Danielson JA, Johanson U (2010) Phylogeny of major intrinsic proteins. In: Jahn TP, Bienert GP

(eds) MIPs and their role in the exchange of metalloids. Landes Biosciences, Austin

Dean RM, Rivers RL, Zeidel ML, Roberts DM (1999) Purification and functional reconstitution of

soybean nodulin 26. An aquaporin with water and glycerol transport properties. Biochemistry

38:347–353

Dordas C, Chrispeels MJ, Brown PH (2000) Permeability ad channel-mediated transport of boric

acid across membrane vesicles isolated from squash roots. Plant Physiol 124:1349–1362

Dowd C, Wilson IW, McFadden H (2004) Gene expression profile changes in cotton root and

hypocotyl tissues in response to infection with Fusarium oxysporum f. sp. vasinfectum. Mol

Plant Microbe Interact 17:654–667

Dynowski M, Schaaf G, Loque D, Moran O, Ludewig U (2008a) Plant plasma membrane water

channels conduct the signalling molecule H2O2. Biochem J 414:53–61

Dynowski M, Mayer M, Moran O, Ludewig U (2008b) Molecular determinants of ammonia and

urea conductance in plant aquaporin homologs. FEBS Lett 582:2458–2462

Fang X, Yang B, Matthay MA, Verkman AS (2002) Evidence against aquaporin-1-dependent CO2

permeability in lung and kidney. J Physiol 542:63–69

Fauteux F, Remus-Borel W, Menzies JG, Belanger RR (2005) Silicon and plant disease resistance

against pathogenic fungi. FEMS Microbiol Lett 249:1–6

Fetter K, Van Wilder V, Moshelion M, Chaumont F (2004) Interactions between plasma mem-

brane aquaporins modulate their water channel activity. Plant Cell 16:215–228

Fischer G, Kosinska-Eriksson U, Aponte-Santamarıa C, Palmgren M, Geijer C, Hedfalk K,

Hohmann S, de Groot BL, Neutze R, Lindkvist-Petersson K (2009) Crystal structure of a

yeast aquaporin at 1.15 angstrom reveals a novel gating mechanism. PLoS Biol 7:e1000130

Fitzpatrick KL, Reid RJ (2009) The involvement of aquaglyceroporins in transport of boron in

barley roots. Plant Cell Environ 32:1357–1365

28 G.P. Bienert and F. Chaumont

Page 27: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Flexas J, Ribas-Carbo M, Diaz-Espejo A, Galmes J, Medrano H (2008) Mesophyll conductance to

CO2: current knowledge and future prospects. Plant Cell Environ 31:602–621

Fortin MG, Morrison NA, Verma DP (1987) Nodulin-26, a peribacteroid membrane nodulin is

expressed independently of the development of the peribacteroid compartment. Nucleic Acids

Res 15:813–824

Fritz R, Bol J, Hebling U, Angerm€uller S, V€olkl A, Fahimi HD, Mueller S (2007) Compartment-

dependent management of H(2)O(2) by peroxisomes. Free Radic Biol Med 42:1119–1129

Fu D, Lu M (2007) The structural basis of water permeation and proton exclusion in aquaporins.

Mol Membr Biol 24:366–374

Fu D, Libson A, Miercke LJ, Weitzman C, Nollert P, Krucinski J, Stroud RM (2000) Structure of a

glycerol-conducting channel and the basis for its selectivity. Science 290:481–486

Fujiyoshi Y, Mitsuoka K, de Groot BL, Philippsen A, Grubm€uller H, Agre P, Engel A (2002)

Structure and function of water channels. Curr Opin Struct Biol 12:509–515

Gazzarrini S, Kang M, Epimashko S, Van Etten JL, Dainty J, Thiel G, Moroni A (2006) Chlorella

virus mt325 encodes water and potassium channels that interact synergistically. Proc Natl

Acad Sci USA 103:5355–5360

Gerbeau P, G€ucl€u J, Ripoche P, Maurel C (1999) Aquaporin Nt-TIPa can account for the high

permeability of tobacco cell vacuolar membrane to small neutral solutes. Plant J 18:577–587

Gonen T, Walz T (2006) The structure of aquaporins. Q Rev Biophys 39:361–396

Gonen T, Cheng Y, Kistler J, Walz T (2004) Aquaporin-0 membrane junctions form upon

proteolytic cleavage. J Mol Biol 342:1337–1345

Guenther JF, Chanmanivone N, Galetovic MP, Wallace IS, Cobb JA, Roberts DM (2003)

Phosphorylation of soybean nodulin 26 on serine 262 enhances water permeability and is

regulated developmentally and by osmotic signals. Plant Cell 15:981–991

Gunnarson E, Axehult G, Baturina G, Zelenin S, Zelenina M, Aperia A (2005) Lead induces

increased water permeability in astrocytes expressing aquaporin 4. Neuroscience 136:105–114

Gupta AB, Sankararamakrishnan R (2009) Genome-wide analysis of major intrinsic proteins in the

tree plant Populus trichocarpa: characterization of XIP subfamily of aquaporins from evolu-

tionary perspective. BMC Plant Biol 9:134

Gustavsson S, Lebrun AS, Norden K, Chaumont F, Johanson U (2005) A novel plant major

intrinsic protein in Physcomitrella patens most similar to bacterial glycerol channels. Plant

Physiol 139:287–295

Hachez C, Moshelion M, Zelazny E, Cavez D, Chaumont F (2006a) Localization and quantifica-

tion of plasma membrane aquaporin expression in maize primary root: a clue to understanding

their role as cellular plumbers. Plant Mol Biol 62:305–323

Hachez C, Zelazny E, Chaumont F (2006b) Modulating the expression of aquaporin genes in planta:

a key to understand their physiological functions? Biochim Biophys Acta 1758:1142–1156

Hachez C, Heinen RB, Draye X, Chaumont F (2008) The expression pattern of plasma membrane

aquaporins in maize leaf highlights their role in hydraulic regulation. Plant Mol Biol

68:337–353

Halestrap AP, Price NT (1999) The proton-linked monocarboxylate transporter (MCT) family:

structure, function and regulation. Biochem J 343:281–299

Hanba YT, Shibasaka M, Hayashi Y, Hayakawa T, Kasamo K, Terashima I, Katsuhara M (2004)

Overexpression of the barley aquaporin HvPIP2;1 increases internal CO(2) conductance and

CO(2) assimilation in the leaves of transgenic rice plants. Plant Cell Physiol 45:521–529

Harries WE, Akhavan D, Miercke LJ, Khademi S, Stroud RM (2004) The channel architecture of

aquaporin 0 at a 2.2-A resolution. Proc Natl Acad Sci USA 101:14045–14050

Heinen RB, Ye Q, Chaumont F (2009) Role of aquaporins in leaf physiology. J Exp Bot

60:2971–2985

Hendriks G, Koudijs M, van Balkom BW, Oorschot V, Klumperman J, Deen PM, van der Sluijs P

(2004) Glycosylation is important for cell surface expression of the water channel aquaporin-2

but is not essential for tetramerization in the endoplasmic reticulum. J Biol Chem

279:2975–2983

Plant Aquaporins 29

Page 28: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Henzler T, Steudle E (2000) Transport and metabolic degradation of hydrogen peroxide in Chara

corallina: model calculations and measurements with the pressure probe suggest transport of

H2O2 across water channels. J Exp Bot 51:2053–2066

Henzler T, Ye Q, Steudle E (2004) Oxidative gating of water channels (aquaporins) in Chara by

hydroxyl radicals. Plant Cell Environ 27:1184–1195

Herrera M, Hong NJ, Garvin JL (2006) Aquaporin-1 transports NO across cell membranes.

Hypertension 48:157–164

Higuchi T, Suga S, Tsuchiya T, Hisada H, Morishima S, Okada Y, Maeshima M (1998) Molecular

cloning, water channel activity and tissue specific expression of two isoforms of radish

vacuolar aquaporin. Plant Cell Physiol 39:905–913

Hill AE, Shachar-Hill B, Shachar-Hill Y (2004) What are aquaporins for? J Membr Biol 197:1–32

Hiroaki Y, Tani K, Kamegawa A, Gyobu N, Nishikawa K, Suzuki H, Walz T, Sasaki S, Mitsuoka

K, Kimura K, Mizoguchi A, Fujiyoshi Y (2006) Implications of the aquaporin-4 structure on

array formation and cell adhesion. J Mol Biol 2355:628–639

Holm LM, Jahn TP, Møller AL, Schjoerring JK, Ferri D, Klaerke DA, Zeuthen T (2005) NH3 and

NH4þ permeability in aquaporin-expressing Xenopus oocytes. Pflugers Arch 450:415–428

Horsefield R, Norden K, Fellert M, Backmark A, T€ornroth-Horsefield S, Terwisscha van Schel-

tinga AC, Kvassman J, Kjellbom P, Johanson U, Neutze R (2008) High-resolution x-ray

structure of human aquaporin 5. Proc Natl Acad Sci USA 105:13327–13332

Husted S, Schjoerring JK (1996) Ammonia Flux between oilseed rape plants and the atmosphere in

response to changes in leaf temperature, light intensity, and air humidity (interactions with leaf

conductance and apoplastic NH4þ and Hþ concentrations). Plant Physiol 112:67–74

Isayenkov SV, Maathuis FJ (2008) The Arabidopsis thaliana aquaglyceroporin AtNIP7;1 is a

pathway for arsenite uptake. FEBS Lett 582:1625–1628

Ishikawa F, Suga S, Uemura T, Sato MH, Maeshima M (2005) Novel type aquaporin SIPs are

mainly localized to the ER membrane and show cell-specific expression in Arabidopsis

thaliana. FEBS Lett 579:5814–5820

Jahn TP, Møller AL, Zeuthen T, Holm LM, Klaerke DA, Mohsin B, K€uhlbrandt W, Schjoerring JK

(2004) Aquaporin homologues in plants and mammals transport ammonia. FEBS Lett 574:31–36

Jauh GY, Phillips TE, Rogers JC (1999) Tonoplast intrinsic protein isoforms as markers for

vacuolar functions. Plant Cell 11:1867–1882

Javot H, Lauvergrat V, Santoni V, Martin-Laurent F, G€ucl€u J, Vinh J, Heyes J, Franck KI,

Sch€affner AR, Bouchez D, Maurel C (2003) Role of a single aquaporin isoform in root

water uptake. Plant Cell 15:509–522

Jeong J, Suh S, Guan C, Tsay YF, Moran N, Oh CJ, An CS, Demchenko KN, Pawlowski K, Lee Y

(2004) A nodule-specific dicarboxylate transporter from alder is a member of the peptide

transporter family. Plant Physiol 134:969–978

Johanson U, Karlsson M, Johansson I, Gustavsson S, Sj€ovall S, Fraysse L, Weig AR, Kjellbom P

(2001) The complete set of genes encoding major intrinsic proteins in arabidopsis provides a

framework for a new nomenclature for major intrinsic proteins in plants. Plant Physiol

126:1358–1369

Johansson I, Karlsson M, Shukla VK, Chrispeels MJ, Larsson C, Kjellbom P (1998) Water

transport activity of the plasma membrane aquaporin PM28A is regulated by phosphorylation.

Plant Cell 10:451–459

Kaldenhoff R, Fischer M (2006) Functional aquaporin diversity in plants. Biochim Biophys Acta

1758:1134–1141

Kaldenhoff R, K€olling A, Richter G (1993) A novel blue light- and abscisic acid-inducible gene of

Arabidopsis thaliana encoding an intrinsic membrane protein. Plant Mol Biol 23:1187–1198

Kamiya T, Fujiwara T (2009) Arabidopsis NIP1;1 transports antimonite and determines antimo-

nite sensitivity. Plant Cell Physiol 50:1977–1981

Kamiya T, Tanaka M, Mitani N, Ma JF, Maeshima M, Fujiwara T (2009) NIP1;1, an aquaporin

homolog, determines the arsenite sensitivity of Arabidopsis thaliana. J Biol Chem

284:2114–2120

30 G.P. Bienert and F. Chaumont

Page 29: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Kato Y, Miwa K, Takano J, Wada M, Fujiwara T (2009) Highly boron deficiency-tolerant plants

generated by enhanced expression of NIP5;1, a boric acid channel. Plant Cell Physiol 50:58–66

Kim YX, Steudle E (2007) Light and turgor affect the water permeability (aquaporins) of

parenchyma cells in the midrib of leaves of Zea mays. J Exp Bot 58:4119–4129

Kjellbom P, Larsson C, Johansson II, Karlsson M, Johanson U (1999) Aquaporins and water

homeostasis in plants. Trends Plant Sci 4:308–314

Klebl F, Wolf M, Sauer N (2003) A defect in the yeast plasma membrane urea transporter Dur3p is

complemented by CpNIP1, a Nod26-like protein from zucchini (Cucurbita pepo L.), and by

Arabidopsis thaliana delta-TIP or gamma-TIP. FEBS Lett 547:69–74

Kojima S, Bohner A, von Wiren N (2006) Molecular mechanisms of urea transport in plants.

J Membr Biol 212:83–91

Kojima S, Bohner A, Gassert B, Yuan L, von Wiren N (2007) AtDUR3 represents the major

transport for high-affinity urea transport across the plasma membrane of nitrogen-deficient

Arabidopsis roots. Plant J 52:30–40Kolla VA, Suhita D, Raghavendra AS (2004) Marked changes in volume of mesophyll protoplasts

of pea (Pisum sativum) on exposure to growth hormones. J Plant Physiol 161:557–562

Law JMS, Fejer SN, Setiadi DH, Chass GA, Viskolcz B (2005) Molecular orbital computations on

lipids: an ab initio exploratory study on the conformations of glycerol and its fluorine

congeners. J Mol Struct 722:79–96

Lee JK, Kozono D, Remis J, Kitagawa Y, Agre P, Stroud RM (2005) Structural basis for conduc-

tance by the archaeal aquaporin AqpM at 1.68 A. Proc Natl Acad Sci USA 102:18932–18937

Leshem Y, Melamed-Book N, Cagnac O, Ronen G, Nishri Y, Solomon M, Cohen G, Levine A

(2006) Suppression of Arabidopsis vesicle-SNARE expression inhibited fusion of H2O2-

containing vesicles with tonoplast and increased salt tolerance. Proc Natl Acad Sci USA

103:18008–18013

Leshem Y, Seri L, Levine A (2007) Induction of phosphatidylinositol 3-kinase-mediated endocy-

tosis by salt stress leads to intracellular production of reactive oxygen species and salt

tolerance. Plant J 51:185–197

Li GW, Peng YH, Yu X, Zhang MH, Cai WM, Sun WN, Su WA (2008) Transport functions

and expression analysis of vacuolar membrane aquaporins in response to various stresses in

rice. J Plant Physiol 165:1879–1888

Li RY, Ago Y, Liu WJ, Mitani N, Feldmann J, McGrath SP, Ma JF, Zhao FJ (2009) The rice

aquaporin Lsi1 mediates uptake of methylated arsenic species. Plant Physiol 150:2071–2080

Lian HL, Yu X, Lane D, SunWN, Tang ZC, SuWA (2006) Upland rice and lowland rice exhibited

different PIP expression under water deficit and ABA treatment. Cell Res 16:651–660

Lin W, Peng Y, Li G, Arora R, Tang Z, Su W, Cai W (2007) Isolation and functional characteri-

zation of PgTIP1, a hormone-autotrophic cells-specific tonoplast aquaporin in ginseng. J Exp

Bot 58:947–956

Liu X, Samouilov A, Lancaster JR Jr, Zweier JL (2002) Nitric oxide uptake by erythrocytes is

primarily limited by extracellular diffusion not membrane resistance. J Biol Chem

277:26194–26199

Liu LH, Ludewig U, Gassert B, Frommer WB, von Wiren N (2003) Urea transport by nitrogen-

regulated tonoplast intrinsic proteins in Arabidopsis. Plant Physiol 133:1220–1228

Lopez F, Bousser A, Sissoeff I, Gaspar M, Lachaise B, Hoarau J, Mahe A (2003) Diurnal

regulation of water transport and aquaporin gene expression in maize roots: contribution of

PIP2 proteins. Plant Cell Physiol 44:1384–1395

Loque D, Ludewig U, Yuan L, von Wiren N (2005) Tonoplast intrinsic proteins AtTIP2;1 and

AtTIP2;3 facilitate NH3 transport into the vacuole. Plant Physiol 137:671–680

Ma JF, Tamai K, Yamaji N, Mitani N, Konishi S, Katsuhara M, Ishiguro M, Murata Y, Yano M

(2006) A silicon transporter in rice. Nature 440:688–691

Ma JF, Yamaji N, Mitani N, Tamai K, Konishi S, Fujiwara T, Katsuhara M, Yano M (2007) An

efflux transporter of silicon in rice. Nature 448:209–212

Plant Aquaporins 31

Page 30: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Ma JF, Yamaji N, Mitani N, Xu XY, Su YH, McGrath SP, Zhao FJ (2008a) Transporters of

arsenite in rice and their role in arsenic accumulation in rice grain. Proc Natl Acad Sci USA

105:9931–9935

Ma N, Xue J, Li Y, Liu X, Dai F, Jia W, Luo Y, Gao J (2008b) Rh-PIP2;1, a rose aquaporin gene, is

involved in ethylene-regulated petal expansion. Plant Physiol 148:894–907

MacRobbie EA (2006a) Osmotic effects on vacuolar ion release in guard cells. Proc Natl Acad Sci

USA 103:1135–1140

MacRobbie EA (2006b) Control of volume and turgor in stomatal guard cells. J Membr Biol

210:131–142

Mahdieh M, Mostajeran A, Horie T, Katsuhara M (2008) Drought stress alters water relations and

expression of PIP-type aquaporin genes in Nicotiana tabacum plants. Plant Cell Physiol

49:801–813

Marjanovi Z, Uwe N, Hampp R (2005) Mycorrhiza formation enhances adaptive response of

hybrid poplar to drought. Ann NY Acad Sci 1048:496–499

Marmagne A, Rouet MA, Ferro M, Rolland N, Alcon C, Joyard J, Garin J, Barbier-Brygoo H,

Ephritikhine G (2004) Identification of new intrinsic proteins in arabidopsis plasma membrane

proteome. Mol Cell Proteomics 3:675–691

Matias AC, Pedroso N, Teodoro N, Marinho HS, Antunes F, Nogueira JM, Herrero E, Cyrne L

(2007) Down-regulation of fatty acid synthase increases the resistance of Saccharomycescerevisiae cells to H2O2. Free Radic Biol Med 43:1458–1465

Maurel C, Reizer J, Schroeder JI, Chrispreels MJ (1993) The vacuolar membrane protein gamma-

TIP creaters water specific chennels in Xenopus oocytes. EMBO J 12:2241–2247

Maurel C, Aquaporins CMJ (2001) A molecular entry into plant water relations. Plant Physiol

125:135–138

Maurel C, Verdoucq L, Luu DT, Santoni V (2008) Annu Rev Plant Biol 59:595–624

Missner A, K€ugler P, Saparov SM, Sommer K, Mathai JC, Zeidel ML, Pohl P (2008) Carbon

dioxide transport through membranes. J Biol Chem 283:25340–25347

Mitani N, Yamaji N, Ma JF (2008) Characterization of substrate specificity of a rice silicon

transporter, Lsi1. Pflugers Arch 456:679–686

Mitani N, Yamaji N, Ma JF (2009) Identification of maize silicon influx transporters. Plant Cell

Physiol 50:5–12

Mittler R (2002) Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci 7:405–410

Mizutani M, Watanabe S, Nakagawa T, Maeshima M (2006) Aquaporin NIP2;1 is mainly

localized to the ER membrane and shows root-specific accumulation in Arabidopsis thaliana.Plant Cell Physiol 47:1420–1426

Moeller HB, Knepper MA, Fenton RA (2009) Serine 269 phosphorylated aquaporin-2 is targeted

to the apical membrane of collecting duct principal cells. Kidney 75:295–303

Mollapour M, Piper PW (2007) Hog1 mitogen-activated protein kinase phosphorylation targets the

yeast Fps1 aquaglyceroporin for endocytosis, thereby rendering cells resistant to acetic acid.

Mol Cell Biol 27:6446–6456

Moshelion M, Becker D, Biela A, Uehlein N, Hedrich R, Otto B, Levi H, Moran N, Kaldenhoff R

(2002) Plasma membrane aquaporins in the motor cells of Samanea saman: diurnal andcircadian regulation. Plant Cell 14:727–739

Mukhopadhyay R, Beitz E (2010) Metalloid transport by aquaglyceroporins: consequences in the

treatment of human diseases. In: Jahn TP, Bienert GP (eds) MIPs and their role in the exchange

of metalloids. Landes Biosciences, Austin

Murata K, Mitsuoka K, Hirai T, Walz T, Agre P, Heymann JB, Engel A, Fujiyoshi Y (2000)

Structural determinants of water permeation through aquaporin-1. Nature 407:599–605

Neill S, Desikan R, Hancock J (2002) Hydrogen peroxide signalling. Curr Opin Plant Biol

5:388–395

Neill S, Barros R, Bright J, Desikan R, Hancock J, Harrison J, Morris P, Ribeiro D,Wilson I (2008)

Nitric oxide, stomatal closure, and abiotic stress. J Exp Bot 59:165–176

32 G.P. Bienert and F. Chaumont

Page 31: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Nemeth-Cahalan KL, Kalman K, Froger A, Hall JE (2007) Zinc modulation of water permeability

reveals that aquaporin 0 functions as a cooperative tetramer. J Gen Physiol 130:457–464

Newby ZE, O’Connell J, Robles-Colmenares Y, Khademi S, Miercke LJ, Stroud RM (2008)

Crystal structure of the aquaglyceroporin PfAQP from the malarial parasite Plasmodiumfalciparum. Nat Struct Mol Biol 15:619–625

Niemietz CM, Tyerman SD (2002) New potent inhibitors of aquaporins: silver and gold com-

pounds inhibit aquaporins of plant and human origin. FEBS Lett 531:443–447

Olbrich A, Hillmer S, Hinz G, Oliviusson P, Robinson DG (2007) Newly formed vacuoles in root

meristems of barley and pea seedlings have characteristics of both protein storage and lytic

vacuoles. Plant Physiol 145:1383–1394

Ouyang LJ, Whelan J, Weaver CD, Roberts DM, Day DA (1991) Protein phosphorylation

stimulates the rate of malate uptake across the peribacteroid membrane of soybean nodules.

FEBS Lett 293:188–190

Ozga JA, van Huizen R, Reinecke DM (2002) Hormone and seed-specific regulation of pea fruit

growth. Plant Physiol 128:1379–1389

Paris N, Stanley CM, Jones RL, Rogers JC (1996) Plant cells contain two functionally distinct

vacuolar compartments. Cell 85:563–572

Park M, Kim SJ, Vitale A, Hwang I (2004) Identification of the protein storage vacuole and protein

targeting to the vacuole in leaf cells of three plant species. Plant Physiol 134:625–639

Peng Y, Lin W, Cai W, Arora R (2007) Overexpression of a Panax ginseng tonoplast aquaporin

alters salt tolerance, drought tolerance and cold acclimation ability in transgenic arabidopsis

plants. Planta 226:729–740

Pettersson N, Filipsson C, Becit E, Brive L, Hohmann S (2005) Aquaporins in yeast and filamen-

tous fungi. Biol Cell 97:487–500

Porquet A, Filella M (2007) Structural evidence of the similarity of Sb(OH)3 and As(OH)3 with

glycerol: implications for their uptake. Chem Res Toxicol 20:1269–1276

Preston GM, Carroll TP, Guggino WB, Agre P (1992) Appearance of water channels in Xenopus

oocytes expressing red cell CHIP28 protein. Science 256:385–387

Preston GM, Jung JS, Guggino WB, Agre P (1993) The mercury-sensitive residue at cysteine 189

in the CHIP28 water channel. J Biol Chem 268:17–20

Rivers RL, Dean RM, Chandy G, Hall JE, Roberts DM, Zeidel ML (1997) Functional analysis of

nodulin 26, an aquaporin in soybean root nodule symbiosomes. J Biol Chem 272:16256–16261

Rojek A, Praetorius J, Frøkjaer J, Nielsen S, Fenton RA (2008) A current view of the mammalian

aquaglyceroporins. Annu Rev Physiol 70:301–327

Romero-Puertas MC, Rodriguez-Serrano M, Corpas FJ, Gomez M, del Rio LA, Sandalio LM

(2004) Cadmium induced subcellular accumulation of O2 and H2O2 in pea leaves. Plant Cell

Environ 27:1122–1134

Sade N, Vinocur BJ, Diber A, Shatil A, Ronen G, Nissan H, Wallach R, Karchi H, Moshelion M

(2009) Improving plant stress tolerance and yieldproduction: is the tonoplast aquaporin

SlTIP2;2 a key to isohydric to anisohydric conversion? New Phytol 18:651–661

Sakurai J, Ishikawa F, Yamaguchi T, Uemura M, Maeshima M (2005) Identification of 33 rice

aquaporin genes and analysis of their expression and function. Plant Cell Physiol 46:1568–1577

Sakurai J, Ahamed A, Murai M, Maeshima M, Uemura M (2008) Tissue and cell-specific

localization of rice aquaporins and their water transport activities. Plant Cell Physiol 49:30–39

Santoni V, Vinh J, Pflieger D, Sommerer N, Maurel C (2003) A proteomic study reveals novel

insights into the diversity of aquaporin forms expressed in the plasma membrane of plant roots.

Biochem J 373:289–926

Santoni V, Verdoucq L, Sommerer N, Vinh J, Pflieger D, Maurel C (2006) Methylation of

aquaporins in plant plasma membrane. Biochem J 400:189–197

Sarda X, Tousch D, Ferrare K, Legrand E, Dupuis JM, Casse-Delbart F, Lamaze T (1997) Two TIP-

like genes encoding aquaporins are expressed in sunflower guard cells. Plant J 12:1103–1111

Sato-Nara K, Demura T, Fukuda H (2004) Expression of photosynthesis-related genes and their

regulation by light during somatic embryogenesis in Daucus carota. Planta 219:23–31

Plant Aquaporins 33

Page 32: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Savage DF, Stroud RM (2007) Structural basis of aquaporin inhibition by mercury. J Mol Biol

368:607–617

Savage DF, Egea PF, Robles-Colmenares Y, O’Connell JD, Stroud RM (2003) Architecture and

selectivity in aquaporins: 2.5 a X-ray structure of aquaporin Z. PLoS Biol 1:E72

Schenk AD, Werten PJ, Scheuring S, de Groot BL, M€uller SA, Stahlberg H, Philippsen A, Engel A(2005) The 4.5 A structure of human AQP2. J Mol Biol 350:278–289

Scheuring S, Buzhynskyy N, Jaroslawski S, Goncalves RP, Hite RK, Walz T (2007) Structural

models of the supramolecularorganization of AQP0 and connexons in junctional microdo-

mains. J Struct Biol 160:385–394

Sch€ussler MD, Alexandersson E, Bienert GP, Kichey T, Laursen KH, Johanson U, Kjellbom P,

Schjoerring JK, Jahn TP (2008) The effects of the loss of TIP1;1 and TIP1;2 aquaporins in

Arabidopsis thaliana. Plant J 56:756–767Secchi F, Lovisolo C, Uehlein N, Kaldenhoff R, Schubert A (2007) Isolation and functional

characterization of three aquaporins from olive (Olea europaea L.). Planta 225:381–392Shimaoka T, Ohnishi M, Sazuka T, Mitsuhashi N, Hara-Nishimura I, Shimazaki K, Maeshima M,

Yokota A, Tomizawa K, Mimura T (2004) Isolation of intact vacuoles and proteomic analysis of

tonoplast from suspension-cultured cells of Arabidopsis thaliana. Plant Cell Physiol 45:672–683Siefritz F, Biela A, Eckert M, Otto B, Uehlein N, Kaldenhoff R (2001) The tobacco plasma

membrane aquaporin NtAQP1. J Exp Bot 52:1953–1957

Siefritz F, Otto B, Bienert GP, van der Krol A, Kaldenhoff R (2004) The plasma membrane

aquaporin NtAQP1 is a key component of the leaf unfolding mechanism in tobacco. Plant J

37:147–155

Smirnoff N (2005) Antioxidants and reactive oxygen species in plants. Blackwell, Oxford

Soria LR, Fanelli E, Altamura N, Svelto M, Marinelli RA, Calamita G (2010) Aquaporin-8-

facilitated mitochondrial ammonia transport. Biochem Biophys Res Commun 393(2):217–221

Stein WD, Danielli JF (1956) Structure and function in red cell permeability. Discuss Faraday Soc

21:238–251

Sui H, Han BG, Lee JK, Walian P, Jap BK (2001) Structural basis of water-specific transport

through the AQP1 water channel. Nature 414:872–878

Sutka M, Alleva K, Parisi M, Amodeo G (2005) Tonoplast vesicles of Beta vulgaris storage rootshow functional aquaporins regulated by protons. Biol Cell 97:837–846

Tajima M, Crane JM, Verkman AS (2010) Aquaporin-4 associations and array dynamics probed

by photobleaching and single-molecule analysis of GFP-AQP4 chimeras. J Biol Chem

285:8163–8170

Tajkhorshid E, Nollert P, Jensen MØ, Miercke LJ, O’Connell J, Stroud RM, Schulten K (2002)

Control of the selectivity of the aquaporin water channel family by global orientational tuning.

Science 296:525–530

Takano J, Noguchi K, Yasumori M, Kobayashi M, Gajdos Z, Miwa K, Hayashi H, Yoneyama T,

Fujiwara T (2002) Arabidopsis boron transporter for xylem loading. Nature 420:337–340

Takano J, Wada M, Ludewig U, Schaaf G, vonWiren N, Fujiwara T (2006) The Arabidopsis major

intrinsic protein NIP5;1 is essential for efficient boron uptake and plant development under

boron limitation. Plant Cell 18:1498–1509

Takano J, Miwa K, Fujiwara T (2008) Boron transport mechanisms: collaboration of channels and

transporters. Trends Plant Sci 13:451–457

Tamas MJ, Luyten K, Sutherland FC, Hernandez A, Albertyn J, Valadi H, Li H, Prior BA, Kilian

SG, Ramos J, Gustafsson L, Thevelein JM, Hohmann S (1999) Fps1p controls the accumula-

tion and release of the compatible solute glycerol in yeast osmoregulation. Mol Microbiol

31:1087–1104

Tanaka M,Wallace IS, Takano J, Roberts DM, Fujiwara T (2008) NIP6;1 is a boric acid channel for

preferential transport of boron to growing shoot tissues in Arabidopsis. Plant Cell 20:2860–2875

Temmei Y, Uchida S, Hoshino D, Kanzawa N, Kuwahara M, Sasaki S, Tsuchiya T (2005) Water

channel activities of Mimosa pudica plasma membrane intrinsic proteins are regulated by

direct interaction and phosphorylation. FEBS Lett 579:4417–4422

34 G.P. Bienert and F. Chaumont

Page 33: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Terashima I, Ono K (2002) Effects of HgCl(2) on CO(2) dependence of leaf photosynthesis:

evidence indicating involvement of aquaporins in CO(2) diffusion across the plasma mem-

brane. Plant Cell Physiol 43:70–78

ThorsenM, Di Y, T€angemo C,Morillas M, Ahmadpour D, Van der Does C,Wagner A, Johansson E,

Boman J, Posas F, Wysocki R, Tamas MJ (2006) The MAPK Hog1p modulates Fps1p-

dependent arsenite uptake and tolerance in yeast. Mol Biol Cell 17:4400–4410

T€ornroth-Horsefield S, Wang Y, Hedfalk K, Johanson U, Karlsson M, Tajkhorshid E, Neutze R,

Kjellbom P (2006) Structural mechanism of plant aquaporin gating. Nature 439:688–694

Tournaire-Roux C, Sutka M, Javot H, Gout E, Gerbeau P, Luu DT, Bligny R, Maurel C (2003)

Cytosolic pH regulates root water transport during anoxic stress through gating of aquaporins.

Nature 425:393–397

Tsukaguchi H, Shayakul C, Berger UV, Mackenzie B, Devidas S, Guggino WB, van Hoek AN,

Hediger MA (1998) Molecular characterization of a broad selectivity neutral solute channel.

J Biol Chem 273:24737–24743

Uehlein N, Lovisolo C, Siefritz F, Kaldenhoff R (2003) The tobacco aquaporin NtAQP1 is a

membrane CO2 pore with physiological functions. Nature 425:734–737

Uehlein N, Fileschi K, Eckert M, Bienert GP, Bertl A, Kaldenhoff R (2007) Arbuscular mycorrhi-

zal symbiosis and plant aquaporin expression. Phytochemistry 68:122–129

Uehlein N, Otto B, Hanson DT, Fischer M, McDowell N, Kaldenhoff R (2008) Function of

Nicotiana tabacum aquaporins as chloroplast gas pores challenges the concept of membrane

CO2 permeability. Plant Cell 20:648–657

Van Wilder V, Miecielica U, Degand H, Derua R, Waelkens E, Chaumont F (2008) Maize plasma

membrane aquaporins belonging to the PIP1 and PIP2 subgroups are in vivo phosphorylated.

Plant Cell Physiol 49:1364–1377

Vandeleur RK, Mayo G, Shelden MC, Gilliham M, Kaiser BN, Tyerman SD (2009) The role of

plasma membrane intrinsic protein aquaporins in water transport through roots: diurnal and

drought stress responses reveal different strategies between isohydric and anisohydric cultivars

of grapevine. Plant Physiol 149:445–460

Vera-Estrella R, Barkla BJ, Bohnert HJ, Pantoja O (2004) Novel regulation of aquaporins during

osmotic stress. Plant Physiol 135:2318–2329

Verkman AS (2009) Knock-out models reveal new aquaporin functions. Handb Exp Pharmacol

190:359–381

Viadiu H, Gonen T, Walz T (2007) Projection map of aquaporin-9 at 7 A resolution. J Mol Biol

367:80–88

Wallace IS, Roberts DM (2004) Homology modeling of representative subfamilies of Arabidopsis

major intrinsic proteins. Classification based on the aromatic/arginine selectivity filter. Plant

Physiol 135:1059–1068

Wallace IS, Roberts DM (2005) Distinct transport selectivity of two structural subclasses of the

nodulin-like intrinsic protein family of plant aquaglyceroporin channels. Biochemistry

44:16826–16834

Wan X, Steudle E, Hartung W (2004) Gating of water channels (aquaporins) in cortical cells of

young corn roots by mechanical stimuli (pressure pulses): effects of ABA and of HgCl2. J Exp

Bot 55:411–422

Wang Y, Tajkhorshid E (2010) Nitric oxide conduction by the brain aquaporin AQP4. Proteins

78:661–670

Weig AR, Jakob C (2000) Functional identification of the glycerol permease activity of Arabi-

dopsis thaliana NLM1 and NLM2 proteins by heterologous expression in Saccharomycescerevisiae. FEBS Lett 481:293–298

Whiteman SA, N€uhse TS, Ashford DA, Sanders D, Maathuis FJ (2008) A proteomic and

phosphoproteomic analysis of Oryza sativa plasma membrane and vacuolar membrane. Plant

J 56:146–156

Wi SG, Chung BY, Kim JH, Lee KS, Kim JS (2006) Deposition pattern of hydrogen peroxide in

the leaf sheaths of rice under salt stress. Biol Plant 50:469–472

Plant Aquaporins 35

Page 34: [Signaling and Communication in Plants] Transporters and Pumps in Plant Signaling Volume 7 || Plant Aquaporins: Roles in Water Homeostasis, Nutrition, and Signaling Processes

Wienkoop S, Saalbach G (2003) Proteome analysis. Novel proteins identified at the peribacteroid

membrane from Lotus japonicus root nodules. Plant Physiol 131:1080–1090

Xu XY, McGrath SP, Meharg AA, Zhao FJ (2008) Growing rice aerobically markedly decreases

arsenic accumulation. Environ Sci Technol 42:5574–5579

Yamaji N, Mitatni N, Ma JF (2008) A transporter regulating silicon distribution in rice shoots.

Plant Cell 20:1381–1389

Yang HC, Cheng J, Finan TM, Rosen BP, Bhattacharjee H (2005) Novel pathway for arsenic

detoxification in the legume symbiont Sinorhizobium meliloti. J Bacteriol 187:6991–6997Yasui M, Hazama A, Kwon TH, Nielsen S, Guggino WB, Agre P (1999) Rapid gating and anion

permeability of an intracellular aquaporin. Nature 402:184–187

Ye Q, Steudle E (2006) Oxidative gating of water channels (aquaporins) in corn roots. Plant Cell

Environ 29:459–470

Ye Q, Wiera B, Steudle E (2004) A cohesion/tension mechanism explains the gating of water

channels (aquaporins) in Chara internodes by high concentration. J Exp Bot 55:449–461

Yu X, Peng YH, Zhang MH, Shao YJ, Su WA, Tang ZC (2006) Water relations and an expression

analysis of plasma membrane intrinsic proteins in sensitive and tolerant rice during chilling and

recovery. Cell Res 16:599–608

Zardoya R (2005) Phylogeny and evolution of the major intrinsic protein family. Biol Cell

97:397–414

Zelazny E, Borst JW, Muylaert M, Batoko H, Hemminga MA, Chaumont F (2007) FRET imaging

in living maize cells reveals that plasma membrane aquaporins interact to regulate their

subcellular localization. Proc Natl Acad Sci USA 104:12359–12364

Zelenina M, Bondar AA, Zelenin S, Aperia A (2003) Nickel and extracellular acidification inhibit

the water permeability of human aquaporin-3 in lung epithelial cells. J Biol Chem

278:30037–30043

Zeuthen T, Klaerke DA (1999) Transport of water and glycerol in aquaporin 3 is gated by H(þ).

J Biol Chem 274:21631–21636

Zhao FJ, Ago Y, Mitani N, Li RY, Su YH, Yamaji N, McGrath SP, Ma JF (2010) The role of the

rice aquaporin Lsi1 in arsenite efflux from roots. New Phytol 186(2):392–399

36 G.P. Bienert and F. Chaumont


Recommended