ISSN 0006-2979, Biochemistry (Moscow), 2025, Vol. 90, Suppl. 1, pp. S1-S35 © Pleiades Publishing, Ltd., 2025.
Russian Text © The Author(s), 2025, published in Uspekhi Biologicheskoi Khimii, 2025, Vol. 65, pp. 3-54.
S1
REVIEW
The Big, Mysterious World of Plant 14-3-3 Proteins
Ilya A. Sedlov
1,2
and Nikolai N. Sluchanko
1,a
*
1
Bach Institute of Biochemistry, Federal Research Center of Biotechnology, Russian Academy of Sciences,
119071 Moscow, Russia
2
Department of Biochemistry, Faculty of Biology, Lomonosov Moscow State University,
119991 Moscow, Russia
a
e-mail: nikolai.sluchanko@mail.ru
Received August 28, 2024
Revised September 6, 2024
Accepted September 20, 2024
Abstract14-3-3 is a family of small regulatory proteins found exclusively in eukaryotic organisms. They
selectively bind to phosphorylated molecules of partner proteins and regulate their functions. 14-3-3 pro-
teins were first characterized in the mammalian brain approximately 60 years ago and then found in plants,
30 years later. The multifunctionality of 14-3-3 proteins is exemplified by their involvement in coordination
of protein kinase cascades in animal brain and regulation of flowering, growth, metabolism, and immunity in
plants. Despite extensive studies of this diverse and complex world of plant 14-3-3 proteins, our understanding
of functions of these enigmatic molecules is fragmentary and unsystematic. The results of studies are often
contradictory and many questions remain unanswered, including biochemical properties of 14-3-3 isoforms,
structure of protein–protein complexes, and direct mechanisms by which 14-3-3 proteins influence the func-
tions of their partners in plants. Although many plant genes coding for 14-3-3 proteins have been identified,
the isoforms for in vivo and in vitro studies are often selected at random. This rather limited approach is
partly due to an exceptionally large number and variety of 14-3-3 homologs in plants and erroneous a priori
assumptions on the equivalence of certain isoforms. The accumulated results provide an extensive but rather
fragmentary picture, which poses serious challenges for making global generalizations. This review is aimed to
demonstrate the diversity and scope of studies of the functions of plant 14-3-3 proteins, as well as to identify
areas that require further systematic investigation and close scientific attention.
DOI: 10.1134/S0006297924603319
Keywords: 14-3-3 protein, plant biochemistry, phosphorylation, protein–protein interactions, enzyme regulation,
signaling pathways
* To whom correspondence should be addressed.
INTRODUCTION
14-3-3 proteins are small acidic dimeric proteins
found only in eukaryotes. They lack enzymatic activ-
ity, but regulate functions of other proteins via bind-
ing to specific phosphorylated sites in partner proteins
and altering their functional properties.
Mammalian 14-3-3 proteins were first described
about 60 years ago; among them, human 14-3-3 pro-
teins have been studied in most detail. The decades
of research have led to the understanding that 14-3-3
proteins are involved in the regulation of many funda-
mental cellular processes, such as intracellular signal-
ing, gene expression, cell cycle control, and cell death.
14-3-3 proteins also play an important role in human
and animal diseases, e.g., viral infections and cancer.
Plant 14-3-3 proteins were discovered 30 years lat-
er than their mammalian homologs. Because plants are
sessile organisms, they must regulate many biochemi-
cal processes in order to adapt to constantly changing
environmental conditions, so regulatory molecules are
of particular importance to them. This is why 14-3-3
regulatory proteins are essential in crucial processes
of plant life, such as response to phytohormones, flow-
ering, growth, mineral nutrition, and plant immunity.
Despite the importance of plant 14-3-3 proteins,
there are still many unexplored issues, or “blank
spots”. Thus, the structure and biochemical properties
SEDLOV, SLUCHANKOS2
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
of plant 14-3-3 proteins have been studied much less
compared to their mammalian counterparts. An insuf-
ficient information on the structure of complexes of
plant 14-3-3 proteins with their interaction partners
explains the lack of understanding of mechanisms
by which 14-3-3 proteins affect the functions of their
partners. In most cases, there is no information on
the details of 14-3-3 interaction with their partners
and effects this binding produces on their activity of
the latter. Many 14-3-3 genes have been identified, but
the isoforms for in vivo and in vitro studies are chosen
almost at random, and the obtained results provide a
very fragmented picture that precludes global gener-
alizations.
This review aims, on one hand, to demonstrate
the diversity and broad scope of studies of the func-
tions of plant 14-3-3 proteins, and, on another hand, to
highlight the fragmentary nature of current concepts
and lack of consistency in the research of plant 14-3-3
proteins.
DIVERSITY AND CLASSIFICATION
OF 14-3-3 PROTEINS
Proteins of the 14-3-3 family are common among
eukaryotes. They were first isolated in the 1960s from
the mammalian brain, where they account for a sig-
nificant fraction (up to 1% [1]) of all proteins. The
name comes from the purification procedure, as these
proteins were found in fraction 14 during chromatog-
raphy on diethylaminoethyl cellulose and produced
spot 3.3 during electrophoresis in a starch gel [2].
Further studies have led to the discovery of seven ho-
mologous 14-3-3 proteins in mammals that have been
designated with the first letters of the Greek alphabet:
α, β, γ, δ, ε, ζ, and η. However, the α and δ isoforms
were later discovered to be the phosphorylated ver-
sions of β and ζ, respectively. We should mention that
defining identified 14-3-3 proteins as “isoforms” was
incorrect, as each 14-3-3 “isoform” is encoded by a
separate gene. Nevertheless, since the term “isoform”
has been commonly accepted for describing 14-3-3
proteins, we will use it in our review. Two specifically
expressed isoforms, σ and τ (or θ), have been discov-
ered later [3]. 14-3-3 proteins have been rediscovered
several times by independent research groups and
in different organisms, which resulted in the emer-
gence of alternative names, for example, Leonardo in
Drosophila fruit fly, BMH1 and BMH2 in Saccharomy-
ces cerevisiae yeast, GRF in Arabidopsis, etc. [3]. Plant
14-3-3 proteins were discovered almost three decades
after their mammalian homologs. In 1992, several pa-
pers on plant 14-3-3 proteins were published at the
same time [4-7]. Perhaps, no other group of organ-
isms has so many alternative names for 14-3-3 pro-
teins as plants, for example, fusicoccin-binding protein
(FCBP) [8], general regulatory factor (GRF) [9], G-box
factor 14-3-3 (GF14) [5, 7], nitrate reductase inhibitor
protein (NIP) [10], tomato fourteen-three-three (TFT)
[11, 12], and rare cold-inducible (RCI) [13]. All these
names belong to members of the same protein family.
They refer to important biological functions of these
proteins and are associated with the history of 14-3-3
protein research. Currently, the names GRF and GF14
are used to classify 14-3-3 proteins from different
plant species. Also thirteen 14-3-3 isoforms from the
model plant Arabidopsis thaliana are named with
Greek letters, by analogy with mammalian proteins,
but starting from the end of the alphabet: omega(ω),
psi  (ψ), chi  (χ), phi  (φ), upsilon  (υ), pi  (π), omicron  (ο),
nu  (ν), mu  (μ), lambda  (λ), kappa  (κ), iota  (ι), and ep-
silon  (ε). The orthologs of A. thaliana isoforms from
other plants can also be designated with Greek letters,
which creates the third system for the naming and
classification of plant 14-3-3 isoforms. Table 1 shows
the correspondence between the three nomenclatures
of 14-3-3 proteins from A. thaliana.
Compared to mammalian species, which have
seven 14-3-3 isoforms, the number of isoforms encod-
ed by plant genomes can vary significantly. For ex-
ample, there are 5 isoforms in strawberry (Fragaria
vesca)  [14], 8 in rice (Oryza sativa)  [15] and cocoa
tree (Theobroma cacao) [16], 13 in tomato (Solanum
lycopersicum)  [11], 18 in soybean (Glycine max) [17],
Table  1. Nomenclatures for 14-3-3 isoforms from
A. thaliana
UniProt
ID
Greek
letters
GRF
nomenclature
GF14
nomenclature
P42643 chi χ GRF1 GF14χ
Q01525 omega ω GRF2 GF14ω
P42644 psi ψ GRF3 GF14ψ
P46077 phi φ GRF4 GF14φ
P42645 upsilon υ GRF5 GF14υ
P48349 lambda λ GRF6 GF14λ
Q96300 nu ν GRF7 GF14ν
P48348 kappa κ GRF8 GF14κ
Q96299 mu μ GRF9 GF14μ
P48347 epsilon ε GRF10 GF14ε
Q9S9Z8 omicron ο GRF11 GF14ο
Q9C5W6 iota ι GRF12 GF14ι
F4IA59 pi π GRF13 GF14π
WORLD OF PLANT 14-3-3 PROTEINS S3
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
26 in tea tree (Camellia sinensis) [16], and 36 in apple
tree (Malus domestica) [18].
The existence of numerous orthologs and par-
alogs, as well as the variability of the total number
of 14-3-3 isoforms in different plant species, are due
to several whole-genome and segmental duplications
that had occurred in plants in the course of evolu-
tion [18-22]. It is believed that an ancient duplica-
tion has led to the emergence of two separate large
14-3-3 phylogenetic lineages called epsilon group and
non-epsilon group. Numerous phylogenetic studies
have confirmed this divergence for 14-3-3 isoforms
from different plant species [11, 14, 16, 18, 19, 22].
There is no established opinion on the time of the
phylogenetic divergence, but it might have occurred
before the emergence of green plants [14]. This de-
marcation of isoforms is fundamental. It can be
traced in all studied species of seed plants (Sperma-
tophyta), as each species in this clade has both ep-
silon and non-epsilon isoforms  [14,  19]. The epsilon
and non-epsilon isoforms differ in the gene structure.
In A.  thaliana, genes coding for the epsilon isoforms
contain 6 to 7 exons separated by 4 to 6 introns, while
genes for the non-epsilon isoforms contain 4 exons
and 3 to 4 introns [23, 24]. Despite the fact that clas-
sification of 14-3-3 paralogs into epsilon and non-ep-
silon groups is commonly recognized, the authors of
some recent studies on the evolution of 14-3-3 proteins
do not divide 14-3-3 isoforms into these two large
groups [24, 25].
In the course of evolution, large phylogenetic
groups have diverged into smaller subgroups. Based
on the analysis of gene sequences from three species
of monocots and nine species of dicots, Ren et al. [18]
identified 11 subgroups (subfamilies) of 14-3-3 genes,
four of which were in the epsilon group and sev-
en – in the non-epsilon group [18]. Some subgroups
in the study were represented exclusively by 14-3-3
genes from the monocots. In another study, which ex-
amined 14-3-3 isoforms from 12 plant species, 4 sub-
groups were identified in the non-epsilon group, and
no subgroups were found in the epsilon group [19].
A very extensive phylogenetic study of the 14-3-3 fam-
ily analyzed isoforms from 46 species of angiosperms,
Fig. 1. Phylogeny and structure of plant 14-3-3 proteins. a) Phylogenetic tree of thirteen 14-3-3 isoforms from A.  thaliana,
constructed with MEGA11 using the maximum likelihood method and rooted using human ζ isoform. Evolutionary dis-
tance scale (0.1) is shown below. b)Structure of 14-3-3 λ dimer from A. thaliana (PDB ID: 8QT5 [26]). The left monomer is
colored according to the ConSurf residue conservation score [27] (default settings), with residues conserved across 14-3-3
species shown in magenta and non-conserved residues shown in cyan, with the corresponding color scale below. AG is
the amphipathic groove (the binding site for the partner protein phosphorylated sequence). The right monomer is colored
green, with the N- and C-termini marked; the alpha-helices are indicated with Roman numerals. c)Amino acid sequences of
phosphorylated motifs in 14-3-3 partner proteins. Phosphorylated residues are shown in red. Positions of residues relative
to the phosphorylated residue (position 0) are shown in gray.
SEDLOV, SLUCHANKOS4
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Table 2. Distribution of plant 14-3-3 isoforms in phylogenetic groups and subgroups (according to Mikhaylova et al. [14]).
Systematic groups
(according to APG IV)
Family Plant species Epsilon
group
Non-epsilon
group
Epsilon
isoforms
Non-
epsilon
isoforms
Total
number
of isoforms
μιεωψκ
Basal angiosperms Amborellales Amborellaceae Amborella trichopoda 1 1 1 1 2 2 10 (6)
Monocots
Zingiberales Musaceae
Musa acuminata
(banana)
611–11 8 2 10
Poales
Bromeliaceae
Ananas comosus
(pineapple)
2– 331 2 7 9
Panicum virgatum
(millet)
3 1 9 - 3 10 13
Sorghum bicolor
(sorghum)
114- 1 5 6
Hordeum vulgare
(barley)
115- 1 6 7
Triticum aestivum
(wheat)
1 2 10 - 1 12 15 (2)
Dicots Superasterids
Ranunculales Ranunculaceae
Aquilegia coerulea
(aquilegia)
121212 4 5 9
Caryophyllales Amaranthaceae
Amaranthus
hypochondriacus
(amaranthus)
33323811
Lamiales Phrymaceae
Mimulus guttatus
(yellow monkeyflower)
121322 4 7 11
Solanales Solanaceae
Solanum tuberosum
(potato)
211422 4 8 12
Solanum lycopersicum
(tomato)
221422 5 8 13
Nicotiana tabacum
(tobacco)
112522 4 9 13
Apiales Apiaceae
Daucus carota
(carrot)
1– 413 1 8 9
WORLD OF PLANT 14-3-3 PROTEINS S5
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Table 2 (cont.)
Systematic groups
(according to APG IV)
Family Plant species Epsilon
group
Non-epsilon
group
Epsilon
isoforms
Non-
epsilon
isoforms
Total
number
of isoforms
μιεωψκ
Dicots Superrosids
Brassicales
Caricaceae Carica papaya (papaya) 2 1 1 1 2 4 3 7
Brassicaceae
Brassica rapa (turnip) 1 3 3 7 4 5 7 16 23
Arabidopsis thaliana 113332 5 8 13
Arabidopsis lyrata 111332 3 8 11
Capsella rubella
(pink shepherd’s purse)
111232 3 7 10
Sapindales Rutaceae Citrus sinensis (orange) 2 1 1 2 1 2 4 5 9
Myrtales Myrtaceae
Eucalyptus grandis
(eucalyptus)
111111 3 3 6
Malvales Malvaceae
Theobroma cacao (cacao) 1 1 1 2 1 1 3 4 7
Gossypium raimondii (cotton) 1 1 2 8 3 2 4 13 17
Cucurbitales Cucurbitaceae
Cucumis sativus
(cucumber)
122212 5 5 10
Fabales Fabaceae
Glycine max (soy) 6 3 - 4 3 2 9 9 18
Trifolium pratense (clover) 4 1 - 1 - 1 5 2 7
Rosales Rosaceae
Prunus persica (peach) 2 1 1 2 - 2 4 4 8
Fragaria vesca (strawberry) 1 - - 1 1 2 1 4 5
Malpighiales
Salicaceae
Populus trichocarpa
(poplar)
322232 7 7 14
Salix purpurea (willow) 2 2 2 2 1 2 6 5 11
Euphorbiaceae
Ricinus communis (castor
bean plant)
111 1 -1 3 2 5
Linaceae Linum usitatissimum (flax) 4 3 1 3 2 4 8 9 17
* APG IV (Angiosperm Phylogeny Group) is a current classification system in botany [28].
** Number of non-classified isoforms is shown in parentheses.
SEDLOV, SLUCHANKOS6
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
including 13 monocots, 32 eudicots, and one represen-
tative of basal angiosperms [14]. The epsilon group
was divided into three subgroups (iota, mu, and ep-
silon), and also three subgroups were distinguished
within the non-epsilon group (omega, psi, and kappa).
This classification was based on the phylogenetic re-
lationships of A.  thaliana isoforms, and the subgroups
were named according to the Greek letter designations
of the most characteristic isoforms. The distribution
of 14-3-3 isoforms in A.  thaliana was the following:
5genes were in the epsilon group [one (ι) in the iota,
one (μ) in the mu, and three (ο, π, ε) in the epsilon
subgroups] and 8 genes were in the non-epsilon group
[three (χ, φ, ω) in the omega, three (υ, ν, ψ) in the
psi, and two (κ, λ) in the kappa subgroups] (Fig. 1a).
As can be noticed from Fig. 1a, each subgroup con-
tained at least one isoform, making A.  thaliana a con-
venient model for a systematic study of 14-3-3 family
representatives. The classification based on the 14-3-3
isoforms of A.  thaliana is complete and the most rep-
resentative, and we will be using it further in this
review.
Numbers of isoforms in subgroups, as well as
the representation of subgroups themselves in plants
may vary significantly. For example, legumes (Faba-
ceae) lack the epsilon subgroup, while the iota and
epsilon subgroups are absent in carrot (Daucus caro-
ta) (Table2). Representatives of the class Monocotyle-
doneae stand out in terms of the isoform distribution
among subgroups. All analyzed representatives of this
class, with the exception of banana (Musa acuminata),
lacked isoforms of the iota subgroup. Plants of the
Poaceae family (grasses) also lack the mu and kappa
subgroups, but the number of isoforms in the psi sub-
group is greatly increased, e.g., up to 10 in wheat (Trit-
icum aestivum) (Table 2). In cereals, the non-epsilon
group contains only ω isoforms. Such distribution has
led to the fact that the psi subgroup has the largest
number of isoforms among the analyzed plants, while
omega is the most widespread subgroup [14].
The high variability in the number and phyloge-
netic affiliation of 14-3-3 paralogs among plant species,
together with the evolution of these proteins through
multiple duplications, have created grounds to assume
that the functions of 14-3-3 isoforms are redundant
and can overlap. Thus, experiments in yeast S.  cerevi-
siae have revealed a high degree of functional similar-
ity between 14-3-3 proteins. S.  cerevisiae contain only
two 14-3-3 genes: BMH1 and BMH2 [29-31]. Knocking
out either of them altered the phenotypes of the re-
sulting mutants only slightly compared to the wild-
type cells, while the double knockout of both genes
was lethal  [29,  31]. However, introduction of plant
14-3-3 gene to the double knockout mutants rescued
a viable phenotype, which indicated interchangeabili-
ty of 14-3-3 proteins even from systematically distant
groups [32]. However, evolutionary analysis of various
angiosperm species revealed the effect of purifying
selection on 14-3-3 proteins, which may be associated
with an acquisition of a specific function or functions
by them [18, 24, 33-35].
Phenotypic analysis of mutant plants deficient by
particular 14-3-3 isoforms can shed light on specific
functions of these proteins in vivo or their redun-
dancy; however, such studies are often unsystemat-
ic and incomplete. Even for the model organism as
A. thaliana, there is no description of phenotypes of
mutants deficient by each of all 13 14-3-3 isoforms,
which hinders elucidation of their in vivo functions.
The studies are often focused on a particular feature
of mutant plants, while neglecting the other traits. In
their extensive analysis of mutations in plant 14-3-3
isoforms, van Kleeff et al. [36] produced single, dou-
ble, triple, and quadruple knockout mutants for 14-3-3
isoforms from the non-epsilon group (λ, κ, ν, υ, φ,
and χ) in A. thaliana. The authors found that the
length of the main root in single and even double mu-
tants did not differ from that in the wild-type plants;
in six triple and three quadruple knockouts, the main
root was shorter than in the wild-type plants  [36].
When three 14-3-3 isoforms from the epsilon group
(ε, μ, and ο) were knocked down simultaneously, the
mutant plants demonstrated serious growth impair-
ments and reduced length of the root and the hypo-
cotyl [37]. Only the knockout of the μ isoform (but
not the ν isoform) resulted in the reduced root length
in A.  thaliana plants grown at constant illumination;
however, both μ- and ν-deficient mutants had shorter
roots when grown under red light [38]. The knockout
of the μ isoform in A.  thaliana reduced the length of
lateral roots, while the overexpression of this isoform
increased it [39]. The transition to flowering under
the long-day conditions was delayed in single μ or ν
knockout mutants [40]. Overexpression of the GF14c
isoform from rice (O. sativa) also delayed flowering,
while its knockout caused an earlier transition to
flowering compared to the wild-type plants [41]. Over-
expression of the Me14-3-3VII isoform from cassava
(Manihot esculenta) in A.  thaliana increased the con-
tent of starch and sugar in the leaves [42]. However,
overexpression of another cassava isoform, Me14-3-3II,
decreased the amount of starch in the mutant plants
[43]. The antisense-mediated knockdown of 14-3-3 ε
and μ resulted in a 2 to 4-fold increase in the amount
of accumulated starch in the mutant plants compared
to the wild-type controls [44]. Overexpression of the
SiGRF1 isoform from the fox millet (Setaria italica) in
A.  thaliana caused an earlier transition to flowering
under the salt stress conditions [25]. Overexpression
of the MdGRF13 isoform from apple tree (M.  domesti-
ca) in A.  thaliana increased plant tolerance to drought
and salt stress [18]. Overexpression of another apple
WORLD OF PLANT 14-3-3 PROTEINS S7
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
isoform (MdGRF6), on the contrary, led to the increase
in the plant sensitivity to exogenous salt, while the
knockdown of this isoform induced plant tolerance to
salt stress [45].
Apparently, mutations in 14-3-3 proteins mostly
affect plant growth (especially root growth), timing of
transition to flowering, tolerance to salt stress, and
starch accumulation. The phenotype of 14-3-3 mutant
plants is determined by the influence of these proteins
on specific physiological and biochemical reactions,
which will be discussed below.
BIOCHEMICAL CHARACTERISTICS
OF PLANT 14-3-3 PROTEINS
Structure of 14-3-3 proteins. Members of the
14-3-3 family are acidic proteins (pI~  4-5) with a mo-
lecular weight of ~30  kDa. In cells, they typically exist
as homo- or heterodimers consisting of the same or
different subunits, respectively. The monomers bind
to each other in an antiparallel fashion through their
N-terminal regions, with the formation of a W-like
structure that has the central symmetry and two
phosphopeptide-binding (amphipathic) grooves, one in
each monomer (Fig.1b). Both plant and animal 14-3-3
proteins consist of 9 α-helices, each located antipar-
allel to a previous one. Helices 1-4 of the monomers
(N-terminal regions) form a surface necessary for di-
merization, while helices 3, 5, 7, and 9 participate in
the formation of the ligand-binding groove. The most
conserved residues are located within the groove,
while the outer surface of the molecule is less con-
served [46, 47] (Fig. 1b).
The C-terminal region of plant 14-3-3 proteins
contains a number of important functional elements,
including a conserved nuclear export signal (NES) in-
herent in mammalian, plant, and fungal 14-3-3 pro-
teins [48, 49]. The role of NES in the functioning of
14-3-3 proteins was shown in Schizosaccharomyces
pombe yeast. Mutation in the NES sequence of Rad24
(14-3-3 family protein) abolished the nuclear export
of phosphorylated Cdc25 which normally signals DNA
damage [50]. Similarly, 14-3-3 was found to regulate
the signaling associated with DNA damage through
COP1 (constitutive photomorphogenic  1) protein, a
ubiquitin ligase of p53 protein. In the case of DNA
damage, COP1 undergoes phosphorylation, interacts
with 14-3-3  σ, and is exported from the nucleus [51].
Mutations in the NES of 14-3-3  σ prevented the export
of COP1 induced by DNA damage [51]. It is believed
that the NES sequence is essential for the regulation of
transcription factors by 14-3-3 proteins through their
retention in the cytoplasm (cytoplasmic sequestering)
and blockade of their entry to the nucleus. Thus, the
binding of 14-3-3  ε to the mitosis-regulating protein
Cdc25 from the clawed frog (Xenopus) prevented
Cdc25 from entering the nucleus and resulted in its
cytoplasmic localization [52].
The C-terminus of 14-3-3 is disordered, non-con-
served, and has different length and amino acid se-
quence in different 14-3-3 isoforms and homologs
from different organisms [53]. Based on the secondary
structure prediction by bioinformatics methods and
data of circular dichroism analysis of the C-terminal
peptide of 14-3-3  ω from A.  thaliana, Shen et al. [54]
suggested that the C-terminal region of this protein
might contain a tenth α-helix. The question on the ex-
istence of this additional C-terminal α-helix remains
unresolved, since the tenth α-helix is absent from any
resolved spatial structure of plant 14-3-3.
It was suggested that the C-terminus of mamma-
lian 14-3-3 isoforms has an autoinhibitory function.
Thus, 14-3-3  ζ lacking the C-terminus exhibited a high-
er affinity for its protein partners Raf-1 and Bad [55].
Using the FRET method, Silhan etal. [56] demonstrat-
ed that the C-terminus is located in the amphipathic
groove and is displaced from it upon phosphopeptide
binding [56]. The autoinhibitory function of the C-ter-
minus in plant 14-3-3 proteins has been proposed
in several studies. Shen et al. [54] showed that the
14-3-3  ω isoform truncated at the C-terminus had a
higher inhibitory activity towards its partner protein
nitrate reductase (NR), thus indirectly indicating the
autoinhibitory role of the C-terminus [54]. However,
Athwal etal. [57] found that the C-terminal truncation
of 14-3-3  ω, on the opposite, significantly reduced the
inhibitory capacity of this protein towards NR  [57].
The autoinhibitory role of the C-terminus was also
observed in the study of the plasma membrane
H
+
-ATPase (AHA1), a well-known partner protein of
14-3-3. Thus, 14-3-3  ω and ε truncated at the C-termi-
nus activated AHA1 more efficiently than the wild-
type isoforms [58]. Deletion of the C-terminus from
the T14-3c isoform of tobacco (Nicotiana tabacum)
increased the protein affinity to sucrose phosphate
synthase (SPS) [59].
Plant 14-3-3 proteins can bind divalent Ca
2+
and
Mg
2+
ions in a region close to the C-terminus. The
binding of calcium and magnesium ions to 14-3-3
and their effect on the protein structure have been
confirmed by various experimental approaches. The
tryptic cleavage patterns of 14-3-3  ω in the presence
and absence of Ca
2+
were different, suggesting confor-
mational rearrangements induced by the binding of
calcium ions [60]. Incubation with radioactive isotope
45
Ca
2+
showed that unlike bovine serum albumin, mem-
brane-immobilized 14-3-3 ω was able to bind Ca
2+
ions
[60]. Equilibrium dialysis experiments demonstrated
that 14-3-3 ω bound calcium at a ratio of one Ca
2+
ion
per one 14-3-3 monomer [60]. The ability of 14-3-3  ω
to bind multiply charged cations was shown using
SEDLOV, SLUCHANKOS8
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
terbium (Tb
3+
)-induced fluorescence, as the appear-
ance of characteristic fluorescence peaks indicated
Tb
3+
binding [61]. The binding of Ca
2+
to the 14-3-3
κ and λ isoforms was demonstrated by thermopho-
resis [62]. It was suggested that the ion-binding site
of 14-3-3 is located in the loop between the helices 8
and 9 [57,  60]. Using manual comparison of sequenc-
es of A. thaliana 14-3-3  ω and calmodulin, Lu et al.
[60] found a similarity between individual amino acid
residues in 14-3-3  ω with the residues in the EF-hand
domain of calmodulin [60]. However, automated bio-
informatics search for A.  thaliana proteins containing
EF-hands performed by Day et al. revealed no EF-hand
domains in 14-3-3 proteins [63]. Using the surface plas-
mon resonance method, it was shown that increasing
concentrations of Ca
2+
and Mg
2+
promoted the binding
of 10 tested 14-3-3 isoforms from A. thaliana, κ, λ,
ν, ω, φ, ψ, υ, ε, and μ) with synthetic phosphopeptides
from the partner proteins nitrate reductase (NR2) and
plasma membrane H
+
-ATPase (AHA2), which indirectly
indicates that the binding of these divalent ions by
14-3-3 can strengthen its interaction with the partner
proteins [64]. It was also noted that the presence of
Ca
2+
and Mg
2+
in the reaction buffer increased the in-
hibitory activity of 14-3-3  ω against NR [57] and its
ability to bind phosphopeptides [61]. On the contrary,
Ca
2+
ions decreased the inhibitory function of 14-3-3  λ
and κ toward the partner protein SOS2 kinase [62].
It should be emphasized that at the moment, there
is no direct experimental evidence of divalent cation
binding in the region between the helices 8 and 9 ob-
tained based on the spatial structure of 14-3-3. Also,
since 14-3-3 proteins are acidic, nonspecific binding of
cations cannot be excluded. Therefore, the questions
about specific binding of Ca
2+
and Mg
2+
, as well as
the identity and location of the cation-binding site in
14-3-3 remain unresolved and controversial.
The properties of 14-3-3 dimers have been studied
in most detail for mammalian 14-3-3 isoforms. Thus,
the dissociation constant (Kd) of human 14-3-3  ζ dimer
measured by several methods was found to be ~5  nM
[65], indicating very strong binding. Human 14-3-3 iso-
forms have also been studied for the monomer pref-
erences upon dimerization. It was found that 14-3-3  ε
tends to form heterodimers with other isoforms [66,
67], while 14-3-3  σ preferentially homodimerizes [68,
69]. A possible explanation lies in the number and
representation of certain amino acid residues and
chemical contacts in the region of dimerization. The
14-3-3 σ monomers are linked in the homodimer by
three pairs of salt bridges (three symmetrical contacts:
Arg19–Glu91, Asp21–Lys87, Lys9–Glu83) and several
additional contacts, which provides a high stability of
the homodimer. According to the hypothesis of Yang
et  al.  [67], 14-3-3  ε monomers form only one pair
of salt bridges in the homodimer (two symmetrical
contacts of Arg19 of one monomer and Glu92 of the
other monomer), while heterodimerization with other
14-3-3 isoforms involves formation of an additional
salt bridge in 14-3-3  ε [e.g., Glu92(ε)–Arg18(ζ) upon in-
teraction with 14-3-3 ζ], which makes 14-3-3 ε heterod-
imerization more preferable [67, 69]. Interestingly, in
some cases, 14-3-3 heterodimers can bind to partner
proteins, as established from the spatial structure of
the human PEAK3 pseudokinase complex with the
14-3-3  ε/β heterodimer [70]. Heterodimers containing
human 14-3-3  ε can efficiently form in vitro under na-
tive conditions via exchange of different homodimer
subunits [71, 72], but the functions of such heterodi-
mers in vivo remain a subject of debate.
Plant 14-3-3 are likely capable of heterodimeriza-
tion as well. 14-3-3 isoforms from A. thaliana were
found to form χ  +  φ/ω/ψ, ω  +  φ/υ/ψ, φ  +  υ/ψ, and ψ  +  υ
heterodimers [73]; the ω isoform was shown to het-
erodimerize with κ and λ isoforms [74]. However, in
the experiments on subunit heterodimerization, the
mixed 14-3-3 isoforms were first denatured in the
presence of guanidine chloride [73] or deoxycholate
[74] and then subjected to renaturation, followed by
analysis of the homo- and heterodimers formed. These
experimental conditions are very far from the native
ones and only indirectly suggested the possibility of
formation of certain dimers in plant cells. Yet, there
is evidence that 14-3-3 heterodimerization can occur
in vivo, although the isoforms involved in this process
remain unknown [73,  75,  76]. It is important to note
that heterodimerization of plant 14-3-3 isoforms has
been studied only for proteins from the non-epsilon
phylogenetic group, while the data on heterodimeriza-
tion of epsilon group isoforms are lacking. At the same
time, an interesting information was obtained by us-
ing the yeast two-hybrid method to study pairwise
dimerization of six 14-3-3 isoforms from the cotton
plant (Gossypium hirsutum), three of which were from
the epsilon group and three – from the non-epsilon
group. The studied isoforms formed only certain het-
erodimers (Gh14-3-3L and Gh14-3-3e, Gh14-3-3L and
Gh14-3-3g, Gh14-3-3a and Gh14-3-3e, Gh14-3-3a and
Gh14-3-3g, Gh14-3-3g and Gh14-3-3h), but not homod-
imers [77]. Verification of these results by direct bio-
chemical methods will be of great interest.
14-3-3 proteins can be phosphorylated, and this
modification can influence dimer dissociation. Phos-
phorylation of Ser58 in the dimer interface of human
14-3-3  ζ resulted in dimer destabilization and dissoci-
ation [78]. In plant 14-3-3 proteins, residues homolo-
gous to mammalian Ser58 can also be phosphorylated
in vivo. The phosphomimetic modification of homol-
ogous Ser62 residue in 14-3-3  ω facilitated dimer
dissociation [79,  80]; this effect was even more pro-
nounced upon introduction of two phosphomimetic
substitutions simultaneously (at Ser62 and Ser67) [79].
WORLD OF PLANT 14-3-3 PROTEINS S9
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Fig.  2. Interaction of 14-3-3 proteins with phosphopeptides. a) Binding of phosphopeptide from BZR1 transcription factor
(motif II) in the amphipathic groove of 14-3-3 ω from A. thaliana (PDB ID: 8QTF [26]). Amino acid residues forming polar
contacts (black dashed line) with the phosphate group and peptide backbone are shown. b) Interaction of 14-3-3 with the
BZR1 phosphopeptide in the absence of fusicoccin (motifII; PDBID: 8QTF [26]); orange, phosphopeptide; red, phosphoser-
ine residue. c) Interaction of 14-3-3 with the PMA2a phosphopeptide and fusicoccin (motif III; PDB ID: 1O9F [88]); blue,
phosphopeptide; red, phosphothreonine residue; pink, fusicoccin molecule. Images were created with the PyMol program.
The phosphomimetic substitution of Ser62 in 14-3-3  ω
disrupted protein heterodimerization preferences and
decreased 14-3-3 affinity for its ligands (peptide difo-
pein and N-terminal fragment of NR) [74]. Interest-
ingly, phosphorylation of 14-3-3  χ at Ser72 (analog of
Ser67 in 14-3-3 ω) decreased its inhibitory activity to-
ward NR [81]. Phosphorylation of residues homologous
to Ser58 in plant 14-3-3 proteins was found to yield
dissociation of calcium-dependent protein kinase  3
(CPK3) from 14-3-3 and the kinase degradation [82].
Therefore, phosphorylation of 14-3-3 proteins at the
dimer interface can significantly alter the biological
functions of these proteins by shifting the dimer/mono-
mer equilibrium. Plant 14-3-3 proteins contain other
phosphorylation sites [46, 83], but the consequences
of their modification and its effect on the structure
and functions of 14-3-3 proteins have been studied
insufficiently.
Principles of 14-3-3 binding to partner proteins.
The binding partners of 14-3-3 proteins are typical-
ly phosphorylated at residues located in structurally
disordered regions [84]. Based on numerous experi-
mental data accumulated through the years of studies,
it was found that 14-3-3 proteins bind phosphoserine
(pS) or phosphothreonine (pT) residues in a specif-
ic amino acid context (consensus motif), and only
in very rare cases interact with nonphosphorylated
partners [85]. There are three types of 14-3-3-bind-
ing phosphorylated motifs: type I – RXXp(S/T)X(P/G),
type II – RX(Y/F)Xp(S/T)X(P/G), and type III (C-termi-
nal)– p(S/T)X
0-2
–COOH, where p(S/T) is phosphoserine/
phosphothreonine and X is any amino acid residue
[48, 86] (Fig.1c). There are data indicating that motifI
in 14-3-3 partner proteins in plants can be slightly dif-
ferent. According to Johnson etal. [87], who conducted
bioinformatics analysis of 14-3-3 recognition sequences
in partner proteins, in many cases, plant motif I is
extended at the N-terminus: LX(R/K)SX(pS/pT)XP [87].
The authors associated this feature with the fact that
many light-dependent processes in plants are regulat-
ed by specific protein kinases that require the pres-
ence of a leucine residue in the beginning of motif I.
However, the number of analyzed 14-3-3 recognition
sites in plants was much smaller than for mammali-
an proteins (14 plant sequences vs. 201 mammalian
sequences) [87], so that the validity of extrapolating
these findings to all partners of plant 14-3-3 proteins
is questionable.
The phosphopeptide binds along its entire length
in the amphipathic groove via polar contacts and hy-
drophobic interactions. The main residues of 14-3-3
that form polar contacts with the phosphopeptide
SEDLOV, SLUCHANKOS10
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
backbone and phosphate group are Lys53, Arg60,
Arg64, Arg133, Tyr134, Asn179, Asn230, and Trp234,
which are highly conserved in plant and animal 14-3-3
proteins [88] (Fig. 2a; residue numbering for Arabi-
dopsis 14-3-3 ω).
14-3-3 isoforms have different affinity for the
partner protein phosphopeptides. Assessment of all 7
human 14-3-3 isoforms for their affinities to phospho-
peptides derived from CTFR  [89], USP8  [90], LRRK2
[91], etc., has shown that γ and η had the highest af-
finities, while σ and ε had the lowest affinities, and
the remaining isoforms had intermediate affinities
for the tested phosphopeptides. The affinity of 14-3-3
isoforms for the phosphopeptides from RSK1, HSPB6,
and papillomavirus E6 protein decreased in the fol-
lowing order: γ>η>ζ>τ>β>ε>σ. The same hierarchy
in the 14-3-3 isoform affinity was observed for large
and diverse samples of phosphopeptides from various
partner proteins [92]. It is quite likely that plant 14-3-3
proteins display a similar hierarchy of affinities for
phosphorylated peptides.
Using the surface plasmon resonance method,
it was demonstrated that 9 (out of 13 known) 14-3-3
isoforms from A. thaliana exhibited different affinities
for the phosphorylated C-terminal peptide of the plas-
ma membrane proton ATPase [47] that decreased in
the following order: φ>χ>ν>ψ>υ>ε>ω>κ>λ. However,
these data contradict the results obtained in the study
on the effect of A. thaliana 14-3-3 isoforms (7 out 13)
on the activity of the same plasma membrane ATPase
[58], which decreased in the order: χ>ω>κ>λ>μ>ε>ο.
The differences in the activity of several 14-3-3 iso-
forms from A. thaliana toward spinach leaf NR was
demonstrated in [93]; in this case, the effect decreased
in the following order: ω>φ>χ>ν [93]. According to
Lambeck et al. [94], who studied the influence of
14-3-3 isoforms from A. thaliana on the NR activity,
the inhibitory effect decreased in the following order:
ω>λ>κ>ν>ψ>χ≫ε (the effect of 14-3-3 ε was so insig-
nificant that the authors were unable to determine
the half-inhibition constant) [94]. In a recent study,
7 isoforms of A. thaliana were examined (μ, ε, and ο
from the epsilon group; λ, ν, ψ, and φ from the non-ep-
silon group) for their binding with the phosphopeptide
from the FD protein (transcription factor and compo-
nent of the florigen activation complex that triggers
flowering). According to the data obtained, the affinity
decreased as follows: ψ>μ>λ>ν>φ>ο>ε [95]. Based on
the above, the data on the affinity of 14-3-3 proteins
to the phosphopeptides of partner proteins are scarce
and contradictory. No correlation was found between
the affinity of a 14-3-3 isoform to the partner proteins
and isoform attribution to the epsilon or non-epsilon
group. Apparently, more systematic studies using all
rather than a few randomly selected isoforms are
required to make confident statements about the ex-
istence and nature of the hierarchy of the binding
affinity of plant 14-3-3 isoforms. It should be noted
that even though such hierarchy has been proven for
7 relatively conserved human 14-3-3 isoforms [92], its
nature remains unclear, especially given that all 7 iso-
forms have the same structure of the phosphorylated
residue-binding pocket in the amphipathic groove. This
suggests that the binding strength may be influenced
by currently unknown allosteric or long-range effects.
In addition to interactions at the primary binding
site located in the amphipathic groove, the partner
proteins can form contacts with other 14-3-3 regions,
as it was observed for the florigen activating com-
plex Hd3a [96] (vide infra). Since the outer regions of
14-3-3 dimers are the least conserved, their interac-
tions may be specific to particular isoforms.
MAIN FUNCTIONS
OF PLANT 14-3-3 PROTEINS
General principles of functioning of 14-3-3
proteins. Similar to their animal and fungal ortho-
logs, plant 14-3-3 proteins lack enzymatic activity and
function by directly interacting with phosphorylated
partner proteins. As a rule, this interaction alters the
activity of the partner protein, thus affecting its bi-
ological function. The mechanisms leading to such
changes can be divided into the following categories.
Enzyme inhibition. In some cases, the binding
of 14-3-3 protein can lead to the enzyme inhibition.
Aclassic example is NR, whose binding to 14-3-3 pre-
sumably causes conformational changes leading to
the suppression of its enzymatic activity [97] (Fig.3a).
Interaction of 14-3-3 with SPS also results in the en-
zyme inhibition [98]. Unfortunately, the lack of struc-
tural data makes it difficult to decipher the mecha-
nism behind the inhibitory effect of 14-3-3.
Enzyme activation. The binding of 14-3-3 to an en-
zyme can also lead to the increase in its catalytic activ-
ity. Thus, the binding of 14-3-3 to the plant H
+
-ATPase
PMA2 results in the removal of the autoinhibitory se-
quence from the catalytic site and enzyme activation
[58]. The spatial structure of the 14-3-3 complex with a
fragment of the plasma membrane ATPase was one of
the first resolved 3D structures. It was found that the
binding of three 14-3-3 dimers induced the formation
of ATPase hexamers  [99] (Fig.  3b). The interaction of
14-3-3 with the cytosolic enzyme glutamine synthase
(GS1) also promoted enzyme activation [100].
Changes in protein location. The binding of 14-3-3
to a partner protein can alter the intracellular loca-
tion of the latter. Thus, the interaction between phos-
phorylated transcription factors BZR1 (BRASSINAZOLE
RESISTANT  1) and BES1 (BRI1 EMS SUPPRESSOR  1)
involved in the intracellular signaling mediated by
WORLD OF PLANT 14-3-3 PROTEINS S11
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Fig.  3. Regulation of functional activity of partner proteins by plant 14-3-3 proteins. a)  NR inhibition by 14-3-3 binding.
b)  Activation of plasma membrane H
+
-ATPase PMA2 by 14-3-3 binding. c)  Changes in the intracellular location (cytoplas-
mic retention) of the transcription factor BZR1 mediated by 14-3-3 proteins. d)  Import of the precursor of Rubisco small
subunit (preSSU) into chloroplasts with the assistance of 14-3-3 and HSP70. e)  Phosphorylation and activation of protein
kinase MAPKKK5 by protein kinase PBL19 in complex with 14-3-3. f) Inhibition of degradation of the transcription factor
WRI after binding to 14-3-3.
brassinosteroids (plant hormones) prevents the entry
of these factors to the nucleus and further regulation
of gene expression [101, 102] (Fig. 3c). Other tran-
scription factors, such as RSG (Repression of Shoot
Growth)  [103] and PIF7 (Phytochrome-Interacting
Factor  7)  [104], also appear to be regulated through
cytoplasmic retention.
Protein transport into chloroplasts and mitochon-
dria. Participation of 14-3-3 in protein transport into
chloroplasts has been suggested already in the early
studies. It was shown that a complex of 14-3-3 with
HSP70 promoted the import of Rubisco small subunit
precursor (preSSU) to chloroplasts [105] (Fig.3d). Also,
14-3-3 was found to interact with the leader sequence
of Arabidopsis photosystem I N-subunit (PSI-N) [106].
According to [107], 14-3-3 promoted the import of at
least 11 chloroplast proteins into chloroplasts. 14-3-3
proteins also participate in protein transport into mi-
tochondria. It is known that the signal sequences of
many proteins imported into mitochondria are phos-
phorylated [108]. In this case, the role of 14-3-3 might
be negative: the binding of MORF3 (Multiple Organel-
lar RNA editing Factor) phosphorylated at the signal
sequence with a complex of 14-3-3 and cytosolic HSP70
slowed down the import of MORF3 into mitochondria
[109]. The details of mechanisms by which 14-3-3 af-
fects the transport of proteins into double-membrane
organelles remain unknown.
Assembly of active protein complexes. 14-3-3 pro-
teins participate in the assembly of functional com-
plexes, e.g., the transcriptional complex with VP1
(Viviparus1) and EmBP1 (Em promoter binding pro-
tein) that activates expression of the Em gene [110].
Formation of complexes with 14-3-3 isoforms can lead
to the protein activation, as was demonstrated for the
activation of MAPKKK5 by a complex of PBL19 kinase
with 14-3-3  λ [111] (Fig. 3e). 14-3-3  κ and λ interact
with the transcription factor PIF3 (Phytochrome-
Interacting Factor  3) and photoactivated phytochrome
phyB and stabilize the complex of these two proteins
SEDLOV, SLUCHANKOS12
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Fig.  4. Mechanisms by which animal 14-3-3 proteins regulate the functions of partner proteins (no similar activities have
been identified for plant 14-3-3 proteins). a)  Protection of PI4KIIIβ from dephosphorylation by 14-3-3  τ binding. b)  Disaggre-
gation of apocytochromec by the chaperone-like activity of 14-3-3  ζ. c)Liquid–liquid phase separation (LLPS) after 14-3-3  ζ
binding to the wild-type tau protein.
in thenucleus, which ultimately results in PIF3 degra-
dation and initiation of photomorphogenesis process-
es [112].
Effect on protein half-life. 14-3-3 proteins can mod-
ulate stability of partner proteins by influencing their
susceptibility to intracellular degradation. For exam-
ple, the binding of 14-3-3 to the key enzyme of eth-
ylene biosynthesis 1-aminocyclopropane-1-carboxyl-
ate synthase (ACS) increases stability of this enzyme,
while 14-3-3 binding to the components of E3 ubiqui-
tin ligase ETO1/EOL, on the contrary, promotes deg-
radation of these proteins [113]. Associated with the
transcription factor WRI1, 14-3-3 prevents its degra-
dation by shielding the binding site for E3 ubiquitin
ligase [114] (Fig. 3f).
Suppression of dephosphorylation. The binding
of 14-3-3 can prevent dephosphorylation of partner
proteins by shielding the phosphate group from phos-
phatases, as demonstrated in studies on human 14-3-3
isoforms. For example, the interaction of 14-3-3 τ with
phosphatidylinositol 4-kinase IIIβ (PI4KIIIβ) increased
the content of phosphorylated active PI4KIIIβ [115]
(Fig.  4a). The binding of 14-3-3 prevented dephos-
phorylation and degradation of the transcription fac-
tor FoxO3 involved in the control of cell proliferation
and apoptosis  [116]. No inhibitors of dephosphoryla-
tion have been found among plant 14-3-3 proteins yet;
however, given their structural and functional similar-
ity to mammalian 14-3-3 proteins, this function can be
expected in plant proteins as well.
Chaperone-like activity. Mammalian 14-3-3 pro-
teins exhibit chaperone-like activity, i.e., prevent ag-
gregation of substrate proteins in an ATP-independent
manner. Mammalian 14-3-3  ζ was found to prevent
aggregation of insulin, alcohol dehydrogenase, and
phosphorylase kinase in vitro [117]; 14-3-3 ζ from
Drosophila prevented aggregation of citrate synthase,
as well as prevented aggregation and even promoted
disaggregation of apocytochrome  c aggregates in vivo
[118] (Fig. 4b). Since this activity does not depend on
phosphorylation of partner proteins, it can be consid-
ered as a “moonlighting” function of 14-3-3 [53]. There
are reasons to believe that plant 14-3-3 proteins also
possess chaperone-like activity, especially considering
a potentially greater susceptibility of plant cells to en-
vironmental changes (e.g., in ambient temperature).
Regulation of liquid-liquid phase separation (LLPS).
Recent studies have demonstrated that mammalian
14-3-3 proteins can modulate LLPS by interacting
with partner proteins. Thus, it was shown that 14-3-3  ζ
binds to unmodified tau protein, promoting formation
of droplets of the tau protein phase during LLPS and
stabilizing them [119] (Fig.4c). It cannot be ruled out
that plant 14-3-3 proteins perform a similar function
in LLPS in plants.
14-3-3 proteins as regulators of plant metabo-
lism. In plants, 14-3-3 proteins regulate primary me-
tabolism by interacting with and controlling the ac-
tivity of key enzymes of nitrogen, carbohydrate, and
sulfur metabolism.
Regulation of nitrogen metabolism. The most stud-
ied protein partner of 14-3-3 in plants is NR, an enzyme
involved in nitrogen metabolism. It is a cytoplasmic
protein that catalyzes the first step of nitrogen assim-
ilation in plants, namely, NADH-dependent conversion
of nitrate to nitrite (Fig.5). Nitrite is reduced in chlo-
roplasts to ammonium ion, which is then incorporat-
ed into amino acids [120, 121]. NR is a large protein;
the molecular weight of its dimer is ~200 kDa [122].
The N-terminal domain of NR contains molybdenum
as a cofactor and is followed by the cofactor-free di-
merization domain, central domain with the heme b5,
and C-terminal domain that contains bound FAD and
has the NADH-binding site [122] (Fig.  5). Because of
its important biological role and dependence on reduc-
ing equivalents (which are actively supplied during the
daytime), NR is tightly regulated by many factors, in-
cluding 14-3-3 proteins [121]. The enzyme is phosphor-
ylated at a serine residue located approximately in
the middle of its polypeptide chain (Ser543 in spinach
NR). The sequence containing this residue corresponds
to the 14-3-3 recognition motif I [61]. Indeed, this mo-
tif is recognized by 14-3-3 proteins [61], leading to NR
inhibition [93, 94, 123]. The binding of 14-3-3 blocks
WORLD OF PLANT 14-3-3 PROTEINS S13
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Fig.  5. Regulation of nitrate reductase (NR) activity by 14-3-3 proteins. Illumination induces formation of reducing and
energy equivalents via photosynthesis. Active NR reduces nitrate ions to nitrites in the cytoplasm. During catalysis, elec-
trons are transferred from NADH to the FAD-containing domain (FAD) and then to the heme  b5-containing domain (b5)
and molybdenum-containing domain (Mo), where nitrate is reduced to nitrite. Nitrite is reduced to ammonium, which is
then incorporated into organic compounds (e.g., amino acids). In the dark, the amount of reducing equivalents decreases
and nitrate reduction stops, so NR should be inhibited. A decrease in the ATP/ADP ratio leads to the SnRK1 activation and
phosphorylation of NR at the 14-3-3 recognition motif located between the Mo and b5 domains (at Ser543 in spinach NR).
The binding of 14-3-3 to this site causes conformational changes in the enzyme, resulting in the spatial separation of the
Mo and b5 domains, which hinders electron transfer between them and leads to NR inactivation.
electron transfer from the heme to the molybdenum
cofactor [97]. Apparently, the binding of 14-3-3 in the
region between the two cofactors causes conforma-
tional rearrangements that move the cofactors away
from each other and make them to adopt a confor-
mation that hinders electron transport [97] (Fig. 5).
Chi et al. [124] provided a more detailed explanation
of this proposed mechanism [124]. They showed that
in NR from A.  thaliana, 14-3-3 binds not only to motif  I
in the proximity of phosphorylated serine, but also to
the conserved acidic N-terminal motif and that com-
plete inhibition of NR requires simultaneous 14-3-3
binding to both sites [124]. Interestingly, the binding
to the acidic N-terminal motif probably occurs outside
the phosphopeptide-binding groove of 14-3-3 [124].
The key serine residue in the 14-3-3-recognizing site
of NR is phosphorylated by the protein kinase SnRK1
(SNF1-related kinase  1) [125]. This enzyme is known to
suppress anabolic and to activate catabolic processes
upon the deficit of energy equivalents in cells [126,
127]. Thus, phosphorylation of NR by SnRK1 under the
energy deficit conditions (in the dark) promotes the
binding of 14-3-3 and conformational rearrangements
leading to the enzyme inhibition. Despite an overall el-
egance of the proposed mechanism, its details remain
poorly understood due to the lack of information on
the structure of the NR complex with 14-3-3 protein.
Cytosolic GS1, a downstream enzyme in the nitro-
gen assimilation pathway, is also regulated by 14-3-3
proteins. GS1 converts glutamic acid to glutamine, re-
sulting in the ammonium incorporation into organic
compounds (Fig. 6a). Finnemann et al. [100] showed
that the binding of 14-3-3 to phosphorylated GS1 in-
creased its catalytic activity in aging rapeseed (Bras-
sica napus) leaves [100]. The phosphorylated residue
in GS1, as well as kinase and phosphatase involved in
the enzyme regulation, remain unknown. Activation
of GS1 in aging leaves is necessary for the nitrogen re-
moval from tissues in the form of organic compounds
[100]. It was also shown that 14-3-3 interacts with
phosphorylated GS1 from Chlamydomonas reinhardtii
[128]. In this case, GS1 is presumably phosphorylat-
ed by the Ca
2+
/calmodulin-dependent protein kinase,
however, neither phosphorylation, nor the binding
of 14-3-3 had any effect on the GS1 activity [128].
Glutamine synthetase GS2 is a GS isoform located in
chloroplasts. Western blotting of native GS2 oligomers
isolated from tomato chloroplasts demonstrated that
SEDLOV, SLUCHANKOS14
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
a b
Fig. 6. Hypothetical scheme of plant metabolism regulation by 14-3-3 proteins. The scheme is based on the most reliable
data (in the opinion of the authors of this review) on the effects of 14-3-3 on the functional activity of partner proteins.
a) Regulation of nitrogen metabolism. b) Regulation of carbohydrate metabolism. Designations: Glc, glucose; Fru, fructose;
UDP-Glc, UDP-glucose; Fru-6-P, fructose 6-phosphate; Sucrose-1-P, sucrose-1-phosphate; GA-3-P, glyceraldehyde 3-phosphate;
3-PGA, 3-phosphoglycerate; 1,3-PGA, 1,3-bisphosphoglycerate; NR, nitrate reductase; GS1, glutamine synthetase 1; SPS, su-
crose-phosphate synthase; CINV1, alkaline invertase 1; GAPN, non-phosphorylating glyceraldehyde 3-phosphate dehydro-
genase; GAPC1, cytosolic glyceraldehyde 3-phosphate dehydrogenase 1.
enzymatically active GS2 octamer contained 14-3-3
[129]. It was suggested that dissociation of 14-3-3 from
the complex results in the loss of GS2 activity [129].
This is an unusual finding, since 14-3-3 typically lo-
calizes to the cytosol or the nucleus. The question of
14-3-3 presence in chloroplasts requires further inves-
tigation. A detailed study on the effect of glutamine
synthetase phosphorylation at various residues on its
interaction with 14-3-3 and enzymatic activity invitro
will provide better understanding of the regulation of
this important enzyme.
Regulation of carbohydrate metabolism. 14-3-3 pro-
teins regulate carbohydrate metabolism at various lev-
els. Several interactome studies have identified dozens
of carbohydrate metabolism enzymes interacting with
14-3-3 proteins [130-133].
SPS catalyzes conversion of fructose-6-phosphate
and UDP-glucose into sucrose-1-phosphate, an imme-
diate precursor of sucrose (Fig.6b). SPS from spinach
(Spinacia oleracea) is regulated by reversible phos-
phorylation at Ser158 that depends on the time of day
and illumination [134]. In the dark, when there is no
photosynthesis and the flow of metabolites decreas-
es, SPS is phosphorylated to inhibit its activity. Expo-
sure to light induces enzyme dephosphorylation and
its activation [134, 135]. It was shown that SPS, like
NR, is phosphorylated by SnRK1 (metabolic regulator
that inhibits anabolic reactions) at Ser158 [125]. Phos-
phorylated SPS binds to 14-3-3 [98, 136]. Using surface
plasmon resonance, Toroser et al. [98] showed that
14-3-3 binds spinach SPS at Ser229 residue. However,
further analysis of 14-3-3 affinity to phosphopeptides
demonstrated that none of the 7 examined A. thaliana
14-3-3 isoforms bound A.  thaliana SPS phosphopeptide
homologous to the Ser229-containing motif from the
spinach SPS; 14-3-3 instead interacted with A. thaliana
SPS phosphopeptide homologous to the Ser158-con-
taining fragment of the spinach enzyme [58]. The stud-
ies using site-directed mutagenesis and yeast two-hy-
brid assay failed to confirm the binding of 14-3-3 to
tobacco SPS at the phosphorylated site homologous to
Ser229 in spinach SPS [59]. Moreover, Ser229 is less
conserved among SPS enzymes from different spe-
cies than Ser158 [137]. These data cast doubts on the
identity of 14-3-3-binding site characterized by Toroser
et al. [98]. It was also shown that 14-3-3 binding
WORLD OF PLANT 14-3-3 PROTEINS S15
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
prevents SPS degradation in vitro [138]. The data on
the effect of 14-3-3 on the SPS activity are controver-
sial. According to Toroser et al., 14-3-3 binding inhib-
ited SPS [98]. However, Moorhead et al. [136] showed
that the phosphopeptide from Raf-1 (mammalian
partner protein of 14-3-3) exhibited a weak inhibito-
ry effect toward SPS, which may indicate activation of
SPS by 14-3-3 proteins [136]. Unfortunately, conflicting
data and lack of information on the structure of 14-3-3
complexes with SPS raise many questions that require
further research.
Alkaline/neutral invertase CINV1 catalyzes the
breakdown of sucrose to glucose and fructose (Fig.6b).
Using a yeast two-hybrid system, it was demonstrated
that CINV1 from A. thaliana binds to 14-3-3 isoforms ε,
κ, λ, υ, φ, χ, ψ, and ω, but not to μ [139]. The binding
is mediated by Ser547, which can be phosphorylat-
ed by calcium-dependent protein kinases [139], and
results in CINV1 activation [139]. Unfortunately, the
information on the structure of CINV1 complexes with
14-3-3 is lacking.
Another carbohydrate metabolism enzyme inter-
acting with 14-3-3 proteins is non-phosphorylating
glyceraldehyde 3-phosphate dehydrogenase (GAPN)
from wheat [140]. This enzyme catalyzes irreversible
oxidation of glyceraldehyde 3-phosphate into 3-phos-
phoglycerate with the reduction of NADP
+
(Fig. 6b).
Itis believed that GAPN is especially important in pro-
viding non-photosynthetic tissues with NADPH [140,
141]. The studies on the effect of phosphopeptides on
the activity of GAPN complex with 14-3-3 revealed
that the interaction between these two proteins caus-
es enzyme inhibition and a 3-fold decrease in the
maximum rate of the GAPN-catalyzed reaction [140].
However, both the 14-3-3 recognition site and the ki-
nase that phosphorylates it remain unknown.
14-3-3 proteins also interact with the cytosolic
glyceraldehyde 3-phosphate dehydrogenase (GAPC1)
from A.  thaliana [138] (Fig. 6b). Most likely, the res-
idue involved in the 14-3-3 binding is Ser126 in the
KVVIpSEP sequence, which is similar to the type  II
recognition motif [58, 142], but the effect of this in-
teraction remains unknown.
Plant 14-3-3 proteins can bind to and influence
the functions of proteins regulating carbohydrate
metabolism. Thus, 14-3-3 has been shown to bind
A.  thaliana trehalose phosphate synthase TPS5 [143],
which catalyzes formation of trehalose 6-phosphate,
a signaling molecule regulating carbohydrate metabo-
lism. The binding is mediated by Ser22 and Thr49 res-
idues located close to the protein N-terminus. In both
cases, the recognition motifs for 14-3-3 are similar to
type motifs, but differ from the canonical ones by
theabsence of proline residue at position  +2  [58,  143].
In vitro, these residues are phosphorylated by SnRK1
[143]. Another important regulatory enzyme of car-
bohydrate metabolism, 6-phosphofructo-2-kinase/fruc-
tose-2,6-bisphosphatase (F2KP), also interacts with
14-3-3. F2KP synthesizes fructose-2,6-bisphosphate, a
signaling metabolite which is believed to direct sug-
ar metabolism toward the synthesis of either sucrose
or starch [144]. Using pull-down assay, Kulma et al.
[145] showed that A. thaliana F2KP specifically bound
14-3-3 in cell extracts, presumably through Ser220 and
Ser303 residues [145]. However, no significant effect
of 14-3-3 on the F2KP activity has been found in vitro,
which might be due to poorly designed experimental
conditions [145].
Many 14-3-3 interaction partners involved in
glycolysis and other carbohydrate metabolism re-
actions have been identified in interactome studies.
These include phosphoenolpyruvate carboxylase, glu-
cose-6-phosphate dehydrogenase [133], phosphoglycer-
ate isomerase, pyruvate kinase, transketolase, sucrose
phosphate synthase [130], triosephosphate isomerase,
enolase, and fructose-bisphosphate aldolase [132].
Unfortunately, this information has not yet been con-
firmed by direct studies of the interaction of these
enzymes with 14-3-3 proteins, which could be an in-
teresting area of future research.
Regulation of sulfur metabolism. Shin et al. [146]
identified a number of enzymes of the sulfur assimi-
lation pathway as 14-3-3 partners [146]. Among them,
two enzymes of cysteine synthesis, o-acetylserine (thiol)
lyase (OASTL) and serine acetyltransferase (SAT), have
been investigated in most detail, and their interaction
with 14-3-3 χ was confirmed by coimmunoprecipita-
tion [146]. However, the details of 14-3-3 interaction
with sulfur metabolism proteins remain enigmatic.
Therefore, there is a large body of evidence indi-
cating that 14-3-3 proteins are involved in the regu-
lation of key enzymes of nitrogen, carbohydrate, and
sulfur metabolism in plants. However, a clear under-
standing of the global role of 14-3-3 in these important
processes is lacking. Moreover, structural and func-
tional details of 14-3-3 interaction with metabolic en-
zymes remain unknown, and reported experimental
data are often contradictory.
Plant 14-3-3 proteins as regulators of mem-
brane transport. 14-3-3 are soluble cytosolic proteins;
however, they can regulate the activity of membrane
transporters and enzymes by binding to them from
the cytosolic side. In a number of studies, plant 14-3-3
proteins have been found in the proximity of the plas-
ma membrane [37, 76, 81, 147].
Regultion of trnsport cross the plsm mem-
brne. One of the most well-characterized protein-pro-
tein interactions of plant 14-3-3 proteins is binding
to the plasma membrane H
+
-ATPase (PMA), or AHA
(Arbidopsis H
+
-ATPase/autoinhibited H
+
-ATPase). PMA
belongs to the family of P-type cation-transporting
ATPases and transfers protons from the cell to the
SEDLOV, SLUCHANKOS16
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
extracellular medium with the expenditure of energy.
It is responsible for the formation of the transmem-
brane proton gradient used as an energy source for
the secondary active transport in plant cells [148].
The C-terminus of PMA contains the autoinhibitory
sequence [149]. PMA binds 14-3-3 at the phosphor-
ylated C-terminal motif YpTV-COOH (type III motif).
The residue that undergoes phosphorylation in PMA2
from Nicotiana plumbaginifolia is Thr948 [150, 151].
This C-terminal motif is a part of the autoinhibitory
domain [152]. The binding of 14-3-3 to PMA abolish-
es autoinhibition and activates the enzyme [58, 151]
(Fig.3b). Activated PMA oligomerizes with the forma-
tion of the active complex containing six PMA subunits
and three 14-3-3 molecules, as was shown by a com-
bination of X-ray crystallography and cryo-electron
microscopy [99] (Fig.3b; Fig. 7a). The activity of PMA
provides a decrease in the apoplast pH and mainte-
nance of the transmembrane proton gradient, which is
necessary for normal plant growth and development,
e.g., elongation of hypocotyl and roots [153], activity of
stomatal cells [99, 154], and pollen tube growth [155].
The binding between 14-3-3 and PMA is a target of
the pathogenic ascomycete fungus Phomopsis amygda-
li (Fusicoccum amygdali). This fungus synthesizes and
secretes the toxin fusicoccin, which greatly increases
the affinity of 14-3-3 for the typeIII motif in PMA by
occupying a free space in the amphipathic groove of
the 14-3-3 complex with the short C-terminal phospho-
peptide [88] (Fig. 2, b and c). Fusicoccin itself binds
rather weakly to 14-3-3 (dissociation constant Kd,
~66 μM), and simultaneous presence of the phospho-
peptide is important for increasing its binding strength
[88, 156]. The mechanism of action of fusicoccin was
elucidated after resolving the spatial structure of its
complex with 14-3-3 and PMA2 phosphopeptide. This
discovery has initiated an entire new direction in the
studies of 14-3-3 proteins. Currently, the structures of
plant 14-3-3 proteins most represented in the Protein
Data Bank are those of N. tabacum 14-3-3 complexes
with fragments of N. plumbaginifolia PMA2 (7 out of
19 structures: 3 complexes with the phosphopeptide
[88, 157] and 4 complexes with a longer C-terminal
fragment of PMA2 [99,  158,  159]) (Table  3). PMA acti-
vation leads to membrane hyperpolarization, opening
of stomata, and uncontrolled transpiration [160]. The
inability to regulate water exchange leads to a patho-
logical phenotype with yellowing and drying leaves
[160]. Notably, one of the first names of plant 14-3-3
was fusicoccin-binding protein (FCBP) due to the dis-
covery of 14-3-3 in the fusicoccin-containing mem-
brane-bound protein complex with PMA [8, 161-163].
Later, it has become clear that fusicoccin stabilizes
not only the binding between 14-3-3 and PMA, but
also other interactions of 14-3-3 with plant and animal
partner proteins through the typeIII motifs [164, 165].
Structural studies have identified a number of small
molecules that can stabilize the interaction of 14-3-3
and PMA2, including cotylenin A [157], pyrrolinone
derivative pyrrolidone 1 [158], dipeptide epibesta-
tin [158], and pyrazole derivative [159]. Analysis of
the obtained spatial structures of protein complexes
with these compounds revealed that the sites of their
binding to 14-3-3 were similar to that of fusicoccin
(Table 3).
Despite generally recognized function of fusicoc-
cin as a PMA activator, there is a growing body of
evidence that this small molecule has other physiolog-
ical effects, such as regulation of potassium ion cur-
rents, enzymes, phytohormones, and protein kinase
Fig.  7. Spatial structures of 14-3-3 complexes with partner
proteins. a) Complex of 14-3-3 isoform C from N. tabacum
(green) with two C-terminal fragments of PMA2 from
N. plumbaginifolia (blue and light blue) and fusicoccin
(red) (PDB ID: 2O98 [99]). b) Complex of 14-3-3 GF14C iso-
form from O. sativa (green) with florigen Hd3a (orange)
and phosphopeptide from OsFD1 transcription factor (pink,
phosphoserine residue is shown in red) (PDBID: 3AXY) [96].
Images were created with PyMol.
WORLD OF PLANT 14-3-3 PROTEINS S17
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Table 3. Spatial structures* of complexes formed by plant 14-3-3 proteins (in reverse chronological order)
PDB ID 14-3-3 protein Partner protein or ligand
Recognition
motif type
Year
released
References
8QTC A. thaliana 14-3-3 ω
phosphopeptide from A. thaliana
transcription factor BZR1
II
2023 [26]
8QTF A. thaliana 14-3-3 ω
phosphopeptide from A. thaliana
transcription factor BZR1
II
8QT5 A. thaliana 14-3-3 λ
phosphopeptide from A. thaliana
transcription factor BZR1
II
8QTT A. thaliana 14-3-3 ω
phosphopeptide from A. thaliana BRI1
kinase inhibitor
8HEW S. tuberosum St14f phosphopeptide from S. tuberosum StFDL1 I 2023 [182]
7XBQ S. tuberosum St14f 2022 [183]
5NWI N. tabacum 14-3-3 C
C-terminal peptide from A. thaliana KAT1
potassium channel
III 2017 [165]5NWJ N. tabacum 14-3-3 C
C-terminal peptide from A. thaliana KAT1
potassium channel
5NWK N. tabacum 14-3-3 C
C-terminal peptide from A. thaliana KAT1
potassium channel and fusicoccin
4DX0 N. tabacum 14-3-3 E
C-terminal peptide from N. plumbaginifolia
ATPase PMA2 and pyrazole derivative
III 2012 [159]
3AXY O. sativa GF14-C
O. sativa florigen Hd3a, peptide from
O. sativa transcription factor OsFD1
I 2011 [96]
3M51 N. tabacum 14-3-3 C
C-terminal peptide from N. plumbaginifolia
ATPase PMA2 and pyrrolidone1
III 2010 [158]
3M50 N. tabacum 14-3-3 C
C-terminal peptide from N. plumbaginifolia
ATPase PMA2 and epibestatin
3E6Y
N. tabacum 14-3-3 C
C-terminal peptide from N. plumbaginifolia
ATPase PMA2 and cotylenin A
III 2009 [157]
2O98 N. tabacum 14-3-3 C
C-terminal peptide from N. plumbaginifolia
ATPase PMA2 and fusicoccin
III 2007 [99]
1O9E N. tabacum 14-3-3 C fusicoccin -
2003 [88]
1O9C N. tabacum 14-3-3 C -
1O9D N. tabacum 14-3-3 C
C-terminal peptide from N. plumbaginifolia
ATPase PMA2
III
1O9F N. tabacum 14-3-3 C
C-terminal peptide from N. plumbaginifolia
ATPase PMA2 and fusicoccin
III
Note. * As of July 2024.
signaling cascades. It was suggested that plants have
an endogenous molecule that has a similar function
as fusicoccin and triggers a coordinated physiological
response involving a number of proteins, i.e., acts as
a phytohormone [166]. Hence, the physiological activ-
ity of fusicoccin remains unclear. It is possible that
new research directions will elucidate it and help to
describe it in detail.
SEDLOV, SLUCHANKOS18
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Fig. 8. The effect of 14-3-3 on the functions of membrane proteins. The binding of 14-3-3 to the cytosolic parts of plasma
membrane integral proteins regulates their activity: the binding of 14-3-3 to PMA2 leads to the enzyme activation; the
interaction with potassium channels KAT1 and GORK promotes ion currents through the channels; the binding to the
aquaporin PIP2;1 maintains its open state. In the vacuole membrane, 14-3-3 interaction with the potassium channel TPK1
decreases the ion current through the channel; the binding of 14-3-3 to SV channels also slows down the cation current
through them; the putative interaction of 14-3-3 with V-ATPase leads to the enzyme activation.
Another function of 14-3-3 proteins is the regu-
lation of potassium ion transport across the plasma
membrane. It has been shown that 14-3-3 enhances
ion currents through the voltage-dependent potassium
channel KAT1 (K
+
channel in A. thaliana) that conducts
potassium into the cell [167, 168] (Fig. 8). Sapona-
ro et al. [165] showed that 14-3-3 interacts with KAT1
through the C-terminal type  III motif YFSpSN-COOH
(phosphorylated residue in A. thaliana is Ser676)
[165]. The authors obtained the spatial structures
of N. tabacum 14-3-3 complex with the phosphopep-
tide from KAT1 (type III motif) in the presence and
absence of fusicoccin [165]. The data suggested that
14-3-3 influences the opening of stomata via two
parallel mechanisms: by acting via PMA ATPase to
generate the proton gradient and hyperpolarize the
membrane in order to promote ion current into the
cells and by affecting KAT1 channels to directly in-
crease K
+
entry into the cells [165]. Both processes ul-
timately lead to the entry of water into stomatal guard
cells and their opening. The role of 14-3-3 proteins in
the regulation of the functioning of stomatal guard
cells is discussed in detail in the review by Cotelle
et al. [147].
Van Kleeff et al. investigated interactions between
14-3-3 and GORK (Guard Cell Outward Rectifier K
+
)
channels from A. thaliana [169] (Fig. 8). These chan-
nels remove potassium ions from cells, thus partici-
pating in the stomata closure and maintenance of K
+
homeostasis in root cells [170]. Using pull-down assay,
the authors showed that GORK channels interact with
14-3-3  λ,  ν, and χ isoforms. However, they failed to
confirm a direct binding of 14-3-3 with GORK channels
in the yeast two-hybrid assay. It was found that 14-3-3
interacts with the calcium-dependent kinase CPK21
and activates its, while CPK21, in turn, phosphorylates
GORK channels at Thr344, Ser518, and Ser649 [169].
The studies of mutant plants deficient by CPK21, 14-3-3
χ, and 14-3-3 φ have shown that these proteins are
required to maintain normal potassium ion currents
through the GORK channels [169].
14-3-3 proteins interact with aquaporins and reg-
ulate their functions. It was found that plant aquapo-
rins from the PIP (Plasma membrane Intrinsic Protein)
WORLD OF PLANT 14-3-3 PROTEINS S19
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
group can bind 14-3-3 proteins [171, 172]. Aquaporins
facilitate water transport across the membrane. Thus,
PIP2;1 from A. thaliana regulates circadian fluid fluxes
in leaves [173]. Using 14-3-3 knockout mutants, it was
demonstrated that 14-3-3 is necessary for maintaining
normal functions of PIP2;1 [172]. It is possible that
PIP2;1 interacts with 14-3-3 through the Ser280 and
Ser283 residues located closer to the C-terminus [172].
It is believed that phosphorylation of these sites has
a regulatory function [173]. These studies again lack
structural information that could shed light on the
mechanism of aquaporin regulation by phosphoryla-
tion and 14-3-3 protein binding.
Regulation of transport across the vacuolar mem-
brane. 14-3-3 proteins also regulate transport through
the vacuolar membrane (tonoplast). In particular,
14-3-3 regulates potassium ion transport to the vacu-
ole lumen through the channels of the TPK/KCO (tan-
dem-pore K
+
channels/K
ir
-like channels) family (Fig.8).
The functions of these channels in plants are poorly
understood. They contain the 14-3-3 recognition motif
I at the N-terminus at the cytosolic side of the mem-
brane (phosphorylated residue in A. thaliana TPK1
is Ser42) and the EF-hand domain at the C-terminus
[174]. The interaction between 14-3-3 λ and A. thaliana
TPK1 was shown using the pull-down assay and the
surface plasmon resonance method [175]. The binding
of three 14-3-3 isoforms (14-3-3 A, B, and C) from bar-
ley (Hordeum vulgare) and HvKCO1 channel (a homo-
log of A. thaliana TPK1) was demonstrated using sur-
face plasmon resonance [176]. The binding of 14-3-3
to TPK1 activates this channel [175]. Interesting data
were obtained when studying TPK1 from H. vulgare.
Two of the three isoforms, 14-3-3 B and C, inhibited
this channel, while 14-3-3 A had no effect on the ion
current [176]. Latz et al. showed that the serine resi-
due in the 14-3-3 recognition motif in A. thaliana TPK1
is phosphorylated by the calcium-dependent protein
kinase CPK3 [177]. The activity TPK1 is important for
maintaining ion homeostasis during plant response to
salt stress [177].
Slow cation (SV) channels are also regulated by
14-3-3 proteins (Fig. 8). These channels conduct Ca
2+
,
K
+
, and Na
+
cations when activated by the increased
intracellular Ca
2+
concentrations [170]. 14-3-3 proteins
slow the current through these channels [175, 178],
although the details of interaction between 14-3-3 and
SV channels remain unknown.
Another protein interacting with 14-3-3 is V-ATPase
[179] (Fig. 8). V-ATPase performs a number of import-
ant functions in the vacuole membrane. By transfer-
ring protons across the membrane, it forms the elec-
trochemical potential, controls turgor pressure, and
stabilizes pH of the cytoplasm [180, 181]. Presumably,
14-3-3 interacts with the A subunit of V-ATPase [179],
but the exact recognition site of 14-3-3 remains un-
known. The amount of 14-3-3 proteins associated with
V-ATPase increases upon illumination with blue light,
as does the V-ATPase activity, but the direct effect of
14-3-3 on the enzyme activity has not been charac-
terized [179].
From the above examples, it is evident that plant
14-3-3 proteins regulated the functions of many mem-
brane proteins, including proton-transporting ATPases,
ion channels, and aquaporins. Many of these 14-3-3
partner proteins are involved in controlling the activi-
ty of stomatal guard cells. Therefore, the opening/clos-
ing of stomata is regulated by 14-3-3 proteins not only
during receiving and transmitting blue light signal via
phototropins, but also directly through the regulation
of ion currents.
14-3-3 proteins as regulators of signaling cas-
cades. Having a sedentary lifestyle, plants cannot
escape from stress factors and have to adjust bio-
chemical characteristics of their cell to environmen-
tal conditions. The effects of various external factors
have led to the formation of a branched signaling net-
work designed to fine-tune the process of biochem-
ical adaptation. By binding phosphorylated residues
in proteins, 14-3-3 proteins can participate in protein
kinase/protein phosphatase signaling cascades or to
add another level of regulation.
Signaling pathways involved in biotic stress re-
sponse. The mitogen-activated protein kinase (MAP)
cascade is one of the best studied in plants. It is a lin-
ear cascade of sequentially phosphorylating kinases:
MAP kinase kinase kinase (MAPKKK) phosphorylates
MAP kinase kinase (MAPKK), which phosphorylates
terminal MAP kinase (MAPK) and activates it [184].
The activation of this signaling pathway can lead to
cell division and differentiation, cell death, and re-
sponse to abiotic and biotic stresses [184]. Oh et al.
[185] found that MAPKKKα from tomato (S. lycoper-
sicum) interacts with the 14-3-3 isoform TFT7 in the
phosphopeptide-binding groove of 14-3-3 [185]. A pos-
sible phosphorylated residue in MAPKKKα is Ser535,
whose amino acid environment corresponds to the
14-3-3 recognition motifII [185]. The presence of 14-3-3
promotes the stability of MAPKKKα protein in cells
and increases the amount of proteins phosphorylated
by this kinase [185]. Overexpression of TFT7 promotes
the pathogen-induced programmed cell death induced
by MAPKKKα [185]. In another study, Oh et  al. [75]
showed that TFT7 binds S. lycopersicum MAPKK iso-
form SlMKK2 (MAPKKKα target) [75]. The putative
recognition site for 14-3-3 is a sequence that includes
phosphorylated Thr33 and corresponds to the type II
motif [75]. In this case, however, the authors did not
find a direct effect of TFT7 on the activity and stability
of SlMKK2 [75]. It was suggested that 14-3-3 has an
adaptor function, since its dimer contains two recog-
nition sites for phosphorylated motifs, one of which
SEDLOV, SLUCHANKOS20
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
can bind SlMKK2, and the other binds MAPKKKα (as
was shown previously), leading to the convergence of
the two kinases and acceleration of signal transduction
[75]. New data to the concept of MAP kinase cascade
regulation by 14-3-3 were added by a recent study by
Dong etal. [111], who demonstrated that 14-3-3 κ and
λ interact with MAPKKK5 and PBL19 from A. thaliana.
PBL19 is a protein kinase of the RLCK family that
phosphorylates the C-terminus of MAPKKK5 and acti-
vates it, but the access of PBL19 to the phosphoryla-
tion site is limited by the autoinhibitory N-terminus of
MAPKKK5. By binding to the C-terminus of MAPKKK5,
14-3-3 λ eliminates this inhibition and allows PBL19 to
phosphorylate and activate MAPKKK5 [111] (Fig. 3e).
Hence, 14-3-3 λ functions as a scaffold that facilitates
MAPKKK5 activation [111], leading to the initiation of
plant immune response against pathogens [111]. How-
ever, as in many other cases, there is no complete un-
derstanding of the MAPK cascade regulation by 14-3-3
proteins due to the lack of structural information on
the relevant complexes.
The 14-3-3 involvement in the regulation of MAPK
cascade explains an important role of 14-3-3 proteins
in the plant immune response to pathogens. It was
found that some effector proteins of pathogenic pro-
karyotes, eukaryotes and viruses can directly interact
with 14-3-3 proteins of host plants. Thus, the coat pro-
tein (CP) of the beet black scorch virus (BBSV, a  (+)RNA
virus) interacts with 14-3-3a of Nicotiana benthamiana
[186]. It was proposed that the recruitment of 14-3-3
by CP leads to a decrease in the activity of another
14-3-3 partner protein, MAPKKKα, which prevents ini-
tiation of cell death and promotes virus replication
[186]. A similar effect was shown for the 14-3-3 part-
ner protein XopQ from the phytopathogenic bacterium
Xanthomonas euvesicatoria. Recruitment of 14-3-3 by
XopQ resulted in the suppression of the MAPKKKα-me-
diated cell death, presumably due to the disruption of
14-3-3 interactions with proteins, including those with
MAPKKKα [187]. Bacteria also secrete other 14-3-3-
interacting effector proteins. For example, Xanthomo-
nas bacteria produce XopX  [188] and XopN1  [189],
while Pseudomonas syringae secretes HopM1  [190]
and HopQ1 [191]. The binding of 14-3-3 with these
effector proteins suppresses plant immune response
to pathogens and facilitates infection. The pathogenic
oomycete Phytophthora palmivora, which causes rot
diseases of many tropical crops, secretes the FIRE
protein into the plant cell. FIRE interacts with 14-3-3
proteins of the host plant via the type I recognition
motif, leading to successful infection; however, the
molecular mechanism of this process remains unclear
[192]. As can be assumed from the presented data,
recruitment of cellular 14-3-3 proteins by binding to
the effector protein is widely used by plant pathogens
to suppress plant immune response, which character-
izes 14-3-3 as an important element in the regulation
of plant immunity [193]. However, both the details of
such interactions and molecular mechanisms of 14-3-3
functioning in plant immune response have not been
clarified so far.
Signaling pathways involved in response to abiot-
ic stress. 14-3-3 proteins participate in the SOS (Salt
Overly Sensitive) signaling cascade, which regulates
plant response to salt stress. The main participants
of the SOS cascade are the serine/threonine protein
kinases SOS2 and PKS5 (SOS2-Like Protein Kinase 5),
Na
+
/H
+
exchanger of the plasma membrane SOS1, and
calcium-binding protein SOS3 [194]. Under normal
conditions, the activities of SOS2 and SOS1 are low. In
the case of salt stress, the concentration of Ca
2+
in cells
increases, which induces SOS3 interaction with SOS2,
SOS2 activation, and phosphorylation of SOS1. Acti-
vated SOS1 pumps sodium ions out of the cell, thus
controlling salt homeostasis [195]. The role of 14-3-3
proteins in the SOS cascade was demonstrated by Yang
etal. [62], who showed that under normal conditions,
PKS5 phosphorylates SOS2 at Ser294, leading to its
interaction with 14-3-3 λ and decrease in the SOS2
activity [62]. However, the amino acid environment
of Ser294 (DGIEGS
294
YVAENV) does not correspond to
the canonical 14-3-3 recognition motifs. Using thermo-
phoresis, the authors showed that 14-3-3  κ and λ are
able to bind Ca
2+
[62]. When the concentration of Ca
2+
increases during the salt stress, 14-3-3 dissociates from
SOS2, resulting in the increase in its kinase activity,
SOS1 activation, and pumping of sodium ions out of
cells [62]. Interestingly, the authors also showed that
14-3-3  λ interacts with PKS5 and inhibits it [62].
Phototropin signaling pathways. Another import-
ant role of 14-3-3 proteins in plants is mediated by
their interaction with the phototropins PHOT1 and
PHOT2. Phototropins are membrane-associated blue-
light receptor protein kinases that initiate intracel-
lular signaling cascade regulating the opening of the
stomata, phototropism, and chloroplast movements
[196,  197]. Phototropins consist of two LOV domains
located at the N-terminus and responsible for captur-
ing light quanta, and the C-terminal kinase domain
that ensures signal transmission [196, 198]. During
photoreceptor activation, a quantum of blue light is
directly captured by flavin mononucleotide (FMN; ab-
sorption maximum, 447  nm) within the LOV domain.
FMN covalently attaches to a cysteine residue, caus-
ing conformational changes leading to phototropin
autophosphorylation and initiation of the signaling
cascade  [196, 198]. Sullivan etal. [199] analyzed inter-
action of PHOT1 and PHOT2 from A. thaliana with six
14-3-3 isoforms, four of which belonged to the non-ep-
silon group (κ, λ, υ, φ), and two were from the epsilon
group (ε, ο). Using far-western blotting, it was demon-
strated that PHOT1 specifically bound the non-epsilon
WORLD OF PLANT 14-3-3 PROTEINS S21
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
isoforms κ, λ, and φ, while the epsilon isoforms (ε, ο)
did not interact with the protein on the membrane
[199]. The authors showed that residues responsible
for the interaction with 14-3-3 (Ser350, Ser376, and
Ser410) were located between the two LOV domains
[199]. At the same time, PHOT2 did not interact with
14-3-3 λ and κ [199]. In a later study, Tseng et al. used
far-western blotting and yeast two-hybrid assay to
demonstrate that PHOT2 interacts with 14-3-3  λ [200].
According to the yeast two-hybrid assay and function-
al studies in mutant plants, the residue responsible for
the interaction with 14-3-3 was Ser747 located in the
kinase domain [200]. The authors demonstrated that
in A. thaliana plants, the effect of PHOT2 on the sto-
mata opening was impaired by mutations in 14-3-3 λ,
but not in 14-3-3  κ [200], indicating a narrow isoform
specificity of 14-3-3 effects. In the absence of stringent
controls, this causes some surprise, since the λ and κ
isoforms are phylogenetically close and have 93% ami-
no acid sequence identity [200]. The results of studies
by Sullivan et al. [199] and Tseng et al. [200] on the
interaction of 14-3-3 and PHOT2 contradict each other
and suggest different 14-3-3 binding sites for such sim-
ilar proteins as PHOT1 and PHOT2. A detailed study
of the interaction of phototropins with 14-3-3 proteins
using structural biology methods could clarify these
issues and explain conflicting observations.
Phytohormone-induced signaling cascades. The sig-
naling cascade of the gaseous phytohormone ethylene
is also regulated by 14-3-3 proteins. In A. thaliana
plants, 14-3-3 proteins interact with ACS, the key
enzyme of ethylene biosynthesis. Yoon  et  al.  [113]
demonstrated that ACS coimmunoprecipitated with
14-3-3  ω, as well as confirmed this interaction in vivo
for four tested 14-3-3 isoforms (ι, ο, κ, φ) using bimo-
lecular fluorescence complementation method. Over-
expression of 14-3-3  ω in cells increased the stability
of ACS, while the treatment of plants with the R18
peptide (which inhibits the binding of 14-3-3 to other
peptides by blocking the amphipathic groove) accel-
erated a decrease in the total ACS content, likely due
to protein degradation. The authors also showed that
14-3-3 ω interacts with ETO1-like (EOL) proteins, com-
ponents of the E3 ubiquitin ligase, which target some
ACS isoforms for degradation  [113,  201]. In a later
study, Catalá etal. [202] also demonstrated the binding
of 14-3-3  ψ (RARE COLD INDUCIBLE 1A, RCI1A) with
ACS using coimmunoprecipitation and bimolecular
fluorescence complementation assays [202]. Howev-
er, the stability of the ACS protein in the presence of
14-3-3  ψ in cells did not increase (as in the case of
14-3-3  ω), but rather decreased. The authors explained
this discrepancy by the specificity of the 14-3-3 ω and
14-3-3  ψ interactions with their partner proteins [202].
In our opinion, these undoubtedly interesting data re-
quire systematization and new studies, in particular,
with the use of structural biology methods, in order
to explain the observed discrepancies.
Plant 14-3-3 proteins as regulators of gene ex-
pression. One of the first names of plant 14-3-3 pro-
tein– G-box-associated factor (GF14)– was due to the
discovery of this protein in a complex with the G-box
element in DNA [7], which suggests participation of
14-3-3 proteins in the regulation of gene transcription.
In A. thaliana, 14-3-3 was found to regulate tran-
scription factors BZR1 and BES1, which are compo-
nents of the brassinosteroid (plant hormone) signaling
pathway. In the absence of brassinosteroids, BZR1 and
BES1 are phosphorylated by the GSK3 family protein
kinase BIN2 at Ser173 [102] and Ser171 [101], respec-
tively. The amino acid environments of the phosphor-
ylated residues represent typeII motifs. Phosphoryla-
tion results in the binding of 14-3-3 (λ and ω isoforms
for BZR1 [26, 102] and λ isoform for BES1 [101]), re-
sulting in the retention of the partner proteins in the
cytoplasm and prevention of their entry to the nucle-
us (Fig. 3c). In the presence of brassinosteroids, the
signaling cascade from the receptor complex of BRI1
(Brassinosteroid Insensitive 1) and BAK1 (BRI1-Asso-
ciated Kinase 1) leads to the inhibition of protein ki-
nase BIN2 by the BSU1 phosphatase [203]. This causes
dephosphorylation of BZR1 and BES1, dissociation of
their complexes with 14-3-3, and import of BZR1 and
BES1 into the nucleus, where they initiate transcrip-
tional programs required for the response to brassi-
nosteroids [101, 102]. Hence, 14-3-3 is involved in the
cytoplasmic sequestration of transcription factors,
possibly due to the presence of the nuclear export
sequence (NES) in the C-terminal region of 14-3-3, as
has been shown for human and yeast 14-3-3 partner
proteins [50, 52].
14-3-3 proteins can be directly involved in the
assembly of transcriptional complexes. It was shown
in O. sativa suspension culture and Zea mays embry-
os, that 14-3-3 is a component of the protein complex
that binds the DNA sequence of the Em1a regulatory
element in the Em gene promoter, whose transcription
is activated by signaling initiated by the phytohor-
mone abscisic acid [110]. The complex also includes
transcription regulators Viviparous1 (VP1) and EmBP1
(b-ZIP family factor) [110]. The association of 14-3-3
with the AtEm1 gene promoter was also shown in
A. thaliana embryonic cell culture [204].
An important role of 14-3-3 proteins is regula-
tion of flowering at the stage of florigen-activating
protein complex formation. The GF14c isoform from
O. sativa interacts in vitro and in vivo with the so-
called florigen, or Hd3a (heading date  3a) protein,
whose homolog in A. thaliana is encoded by the FT
(FLOWERING LOCUST) gene [41,  96]. Florigen is syn-
thesized in the leaves and transported to the apical
meristem of shoots, where it enters the cell cytoplasm
SEDLOV, SLUCHANKOS22
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
and interacts with GF14c. The resulting complex en-
ters the nucleus and binds to the transcription fac-
tor OsFD1 (FD is OsFD1 homolog in A. thaliana) [96].
The florigen-activating complex (FAC) formed by the
GF14c dimer, OsFD1 dimer, and two Hd3a proteins,
activates transcription of OsMADS15 (its homolog in
A. thaliana is called APETALA1, or AP1), thus trigger-
ing the flowering [96]. The spatial structure of a FAC
fragment was solved by X-ray crystallography; it is
the only structure of plant 14-3-3 complex with a full-
length partner protein in the PDB (PDBID: 3AXY; 164
amino acid residues out of 179 resolved in Hd3a struc-
ture) (Fig.7b) [96]. OsFD1 is phosphorylated at Ser192
(type I motif) and bound in the groove of 14-3-3 (as
follows from the structure of 14-3-3 complex with a
OsFD1 phosphopeptide) [96]. In A. thaliana, FD tran-
scription factor is phosphorylated at Thr282 (residue
homologous to Ser192 in OsFD1) by calcium-depen-
dent protein kinases, such as CPK6 and CPK33 [205].
14-3-3 proteins can indirectly regulate gene ex-
pression by affecting the stability of transcription fac-
tors. This regulatory mechanism has been shown for
the transcription factor WRINKLED1 (WRI1), a master
regulator of triacylglyceride biosynthesis [114]. WRI1
likely interacts with 14-3-3 at the site that was also
predicted to bind E3 ubiquitin ligase [114]. Hence,
14-3-3 can protect WRI1 from proteasomal degrada-
tion and promote expression of WRI1-regulated genes
of triglyceride biosynthesis [114] (Fig. 3f).
Guo et al. showed that 14-3-3 (HbGF14a isoform)
from the rubber tree (Hevea brasiliensis) regulates the
activity of the transcription factor HbRZFP1 (H. brasil-
iensis RING zinc finger protein) [206]. HbRZFP1 binds
to the promoter of the gene coding for Hevea rub-
ber transferase (HRT2), an enzyme that adds isoprene
units to the growing rubber polymer, and inhibits its
expression. HbGF14a has been shown to interact with
HbRZFP1 in vivo and in vitro in bimolecular fluores-
cence complementation, yeast two-hybrid, and pull-
down assays [206]. A regulatory mechanism has been
proposed in which the binding of HbGF14a to the
transcription factor HbRZFP1 disrupts HbRZFP1 in-
teraction with the HRT2 gene promoter, resulting in
upregulated expression of this transferase [206]. Sev-
eral studies have also investigated the role of 14-3-3
proteins in rubber biosynthesis. HbGF14c (another
14-3-3 isoform from H.brasiliensis) has been shown to
interact with the small rubber particle protein (SRPP),
which is present in rubber particles and involved in
rubber biosynthesis [207, 208]. However, the details
of this interaction and the corresponding regulatory
mechanism remain unknown.
As can be seen from the above, 14-3-3 proteins
participate in the regulation of many transcription
factors and DNA–protein complexes. However, the ef-
fects and the mechanisms of this regulation are often
poorly understood, while the details of protein–pro-
tein interactions still remain enigmatic. Using mod-
ern methods of structural biology and studying the
processes at the protein level might help in resolving
these issues.
CONCLUSION. “BLANK SPOTS”
IN THE STUDIES OF PLANT 14-3-3 PROTEINS
As can be seen from the data presented in this
review, plant 14-3-3 proteins are involved in the reg-
ulation of many important biochemical and physiolog-
ical processes. The fact of 14-3-3 family involvement
in these processes is beyond any doubt. Hundreds of
research papers and dozens of reviews have been
devoted to the functions of 14-3-3 proteins in plants.
However, despite more than thirty years of research,
many issues about the structure and functions of plant
14-3-3 proteins remain unexplored. The majority of
these fundamental problems arise due to the lack of
systematic approach and, as a consequence, a frag-
mentary nature of the obtained data. 14-3-3 proteins
are a multigene family, and plant species often have
several 14-3-3 isoforms, so to obtain a complete pic-
ture, it is necessary to study either all representatives
of the family or a representative set of these isoforms.
In the first case, the task is extremely labor-intensive
even for plant species with a relatively small number
of isoforms. In the second case, it is unclear which
criteria should be used to select 14-3-3 isoforms for
analysis. Therefore, in most studies, isoforms are cho-
sen at random, which complicates data interpretation
and makes global generalizations impossible. Below,
we present some of the most interesting unresolved
issues – “blank spots” – in the field of plant 14-3-3
research.
Are biochemical properties of plant 14-3-3 ho-
modimers similar or different? One of the funda-
mental questions is related to the biochemical char-
acteristics of 14-3-3 proteins. These proteins are quite
conserved; the pairwise identity of amino acid se-
quences of 14-3-3 isoforms from A. thaliana is 60% on
average and reaches 80-90% in some cases. However,
the question whether their biochemical characteristics
are the same has not been properly addressed. When
we started to investigate this problem, we were sur-
prised to find that 14-3-3 isoforms from the epsilon
and non-epsilon subgroups in A. thaliana differ sig-
nificantly in the stability of formed homodimers, hy-
drophobicity of protein surface, thermal and proteo-
lytic stability, and other properties [209]. Thus, unlike
non-epsilon isoforms, isoforms from the epsilon group
are prone to dissociation into monomers, have a lower
half-transition temperature (the greatest difference in
the half-transition temperature between the epsilon
WORLD OF PLANT 14-3-3 PROTEINS S23
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
and non-epsilon isoforms was over 20°C), and are
more susceptible to limited proteolysis [209]. Even the
homodimers of plant 14-3-3 proteins can significantly
differ from each other. The search for such distinctive
features and biochemical characteristics is of funda-
mental importance, since they can strongly impact
the half-lifetime (stability) and biological functions
of 14-3-3 isoforms in plants. Also, understanding spe-
cific properties of isoforms from different subgroups
can help in phylogenetic analysis and projecting the
properties inherent in a subgroup or in individual
isoforms in some plants onto other less-studied iso-
forms (and in less-studied plants). In the absence of
such information, the growing body of data cannot be
properly systematized and understood.
Which plant 14-3-3 isoforms form heterodi-
mers? Is there a preference for particular isoforms
during heterodimer formation? Heterodimerization
of plant 14-3-3 proteins, both in vitro and in vivo, has
been demonstrated in several studies. It was found
that 13 isoforms of A. thaliana 14-3-3 proteins can
theoretically form 78 different heterodimers, but only
11 of them have been reported to generate heterod-
imers, all of these isoforms belonging to the non-ep-
silon phylogenetic group. Moreover, there is a lack of
information on the ability of plant 14-3-3 proteins to
heterodimerize in their native form, without the use
of denaturation/renaturation. It remains unknown
whether plant 14-3-3 isoforms from the epsilon sub-
group heterodimerize with each other, and whether
epsilon and non-epsilon isoforms can form heterod-
imers. The biological function of heterodimerization
and advantages it provides (if any) remain obscure.
Is there a hierarchy of affinity of plant 14-3-3
proteins for phosphorylated peptides? What is the
reason for the differences in the affinity of 14-3-3 iso-
forms? It has been shown that regardless of the type
of the partner protein, the hierarchy in the affinity
of animal 14-3-3 isoforms remains more or less the
same [92]. It is unknown whether this is true for plant
14-3-3 proteins. Several studies have shown that 14-3-3
isoforms from A. thaliana bind phosphopeptides (or
partner proteins) with different affinity. A systematic
approach is needed to figure out the patterns in the
affinity of plant 14-3-3 proteins, since none of the
studies have examined the full set of isoforms even
for such model plant as A. thaliana. The maximum
number of isoforms analyzed in one study was 9 (out
of 13). The reason for the difference in the affinity
between the 14-3-3 proteins still has to be elucidated.
The residues exposed in the ligand-binding groove of
14-3-3 and responsible for the binding of phosphopep-
tides are fully conserved, so it remains obscure what
provides the differences in the binding affinity from
the structural point of view. This question is relevant
for both plant and mammalian 14-3-3 proteins.
How widespread and universal is the regula-
tion of plant 14-3-3 proteins by phosphorylation?
What are the functional effects of phosphoryla-
tion? 14-3-3 proteins can undergo phosphorylation,
which regulates their binding to the partner proteins
and the 14-3-3 dimer structure [74, 82]. Large-scale
phosphoproteomic studies have identified dozens of
residues capable of being phosphorylated in plant
14-3-3 proteins [83, 210]. However, only two sites locat-
ed at the 14-3-3 dimer interface have been functionally
characterized. It also remains unclear to what extent
the regulation by phosphorylation is conserved and
universal across plant 14-3-3 isoforms.
Are the functions of plant 14-3-3 isoforms spe-
cific or redundant? This question has already been
raised in several reviews on plant 14-3-3 proteins [46,
211, 212]. Does each 14-3-3 isoform bind specific part-
ner proteins or can all 14-3-3 isoforms interact with
all formally suitable partner proteins and compensate
for each others functions? A clear answer to this ques-
tion is currently lacking. For example, 14-3-3  κ and λ
isoforms from A. thaliana have appeared as a result
of a relatively recent duplication of a single gene [14].
Their amino acid sequences are as high as 93% identi-
cal, which is the highest identity of the primary struc-
tures of 14-3-3 isoforms in A. thaliana. For such closely
related isoforms, one would expect a strong overlap
of functions and the least degree of subfunctionaliza-
tion. This is consistent with some experimental data
characterizing the functions of these isoforms as re-
dundant. Thus, single mutations in either 14-3-3  λ and
κ had no effect on the flagellin-induced expression
of marker genes [111]. Both single mutants, as well
as the double mutants by 14-3-3 λ and κ, did not dif-
fer from the wild-type plants in the root length [36].
However, other studies have demonstrated distinct
properties and effects of these highly related isoforms:
14-3-3  λ (but not κ) bound RPW8.2 protein in immu-
noprecipitation experiments [213]; 14-3-3  λ (but not  κ)
was shown to play a role in the PHOT2-mediated
opening of stomata in vivo [200]; both 14-3-3  λ and
κ interacted with the PHOT1, but 14-3-3  λ bound the
protein several-fold more strongly than 14-3-3 κ [199];
the affinity of 14-3-3  λ for various targets was consis-
tently higher than that of 14-3-3  κ  [58]. Thus, even
very similar 14-3-3 isoforms can have both specific
and overlapping functions. Apparently, there is no
general rule about redundancy or specificity of plant
14-3-3 proteins, and the extent of manifestation of the
functional effects of particular isoforms depends on
specific protein–protein interactions.
Lack of information on plant 14-3-3 partner
proteins. Another serious problem is the lack of in-
formation on the interactions of 14-3-3 isoforms with
partner proteins, in particular, on the 14-3-3 binding
site, functional consequences of such interactions,
SEDLOV, SLUCHANKOS24
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
and spatial structures of complexes formed. Solving a
spatial structure naturally helps to identify the mech-
anism of regulation of partner proteins by 14-3-3 iso-
forms. At present, only two structures of plant 14-3-3
complexes with a partner protein (florigen Hd3a from
O. sativa) or a large fragment (C-terminal fragment of
the plasma membrane H
+
-ATPase PMA2 from N.plum-
baginifolia) have been determined [96, 99] (Fig.  7,  a
andb). The details of how 14-3-3 proteins affect their
interaction partners are either unknown or incom-
plete, which leads to ambiguity in the interpretation
of experimental data and contradictions. For example,
neither the phosphorylated residue (Ser229) in SPS
from S. oleracea originally identified by Toroser etal.
[98], nor the inhibitory effect of 14-3-3 upon binding
to the enzyme have been confirmed in later studies.
The regulatory mechanism has been most fully de-
scribed for the two classical 14-3-3 partner proteins
NR [124] and PMA [99], but for many other proteins,
the details of interaction with 14-3-3 remain unknown.
Thus, there are no data on the 14-3-3 binding sites in
GAPN, ACS, GS1, and transcription factors WRI1, VP1,
and EmBP1, as well as on the kinases that phosphor-
ylate them. Based on the general concepts of 14-3-3
interaction with partner proteins, it might be possible
to fully describe the regulatory mechanism if the fol-
lowing information is revealed: (i)  14-3-3 recognition
site in the partner protein; (ii)  kinase that phosphory-
lates it; (iii)  effect of 14-3-3 binding on the activity and
conformation of the partner protein in vitro; (iv) func-
tional consequences of this interaction in vivo. Only
by answering all these questions, researchers will be
able to resolve existing contradictions and clarify the
mechanisms of 14-3-3-regulated processes in plants.
Is yeast two-hybrid assay applicable to study-
ing 14-3-3 interactions with partner proteins? Yeast
two-hybrid assay has been repeatedly used in the stud-
ies of 14-3-3 proteins. The screening of cDNA libraries
has revealed tens and hundreds of protein–protein
interactions [171, 214]. However, the method may not
be suitable for testing specific binary interactions due
to specific features of 14-3-3 interaction with partner
proteins. The binding of 14-3-3 requires phosphory-
lation of the partner protein at a specific motif by a
specific kinase, which might exist in plants only. The
formed protein complexes might not localize to the
nucleus, for example, if the partner protein is located
in the membrane or near it. These features complicate
the use of yeast two-hybrid assay and limit its appli-
cations in the studies of protein–protein interactions
of 14-3-3 isoforms. For example, the conclusion about
the absence of direct binding between 14-3-3 and
GORK channels based on the negative results of yeast
two-hybrid assay is questionable, as in the same study,
specific binding of these two proteins was demon-
strated by the pull-down assay [169]. Therefore, yeast
two-hybrid assay should be used with great caution
and, preferably, in combination with other methods
when studying protein interactions involving 14-3-3
proteins.
Can plant 14-3-3 proteins regulate the function-
ing of partner proteins through the chaperone-like
activity or by influencing the LLPS? A number of
mechanisms by which the binding to 14-3-3 changes
the biological function of partner proteins have been
described for both plant and mammalian 14-3-3 pro-
teins (Fig.3,a-e). However, the chaperone-like activity
and effect on LLPS have been found only for mamma-
lian 14-3-3 proteins [117-119] (Fig. 4, b and c). Since
plant 14-3-3 proteins are structurally similar to their
mammalian homologs and have the same biological
significance, it is reasonable to assume that similar
mechanisms can be found in plants as well. In plants,
LLPS has been described for several important phys-
iological processes, including formation of molecular
condensates in the nucleus during regulation of flow-
ering and initiation of transcriptional response in the
phytochrome phyB signaling [215]. 14-3-3 proteins are
directly involved in the latter two processes and may
play a role in the occurring LLPS events.
Are 14-3-3 proteins present in chloroplasts and
mitochondria and how do they get there? 14-3-3
proteins are typically located in the cytosol and nu-
cleus and lack a signal sequence for their import into
chloroplasts [216]. However, several early studies have
shown that plant 14-3-3 proteins can localize to the
chloroplast stroma [44,  106]. Moreover, it has been
shown that 14-3-3 isoforms can interact with and reg-
ulate the activity of several chloroplast proteins, such
as GS2 [129], starch synthase [44], and chloroplast
and mitochondrial ATP synthases [217]. It remains un-
known how 14-3-3 proteins get into the double-mem-
brane organelles and what functions they perform
there. These issues, addressed mainly in early works,
need to be reconsidered and investigated using mod-
ern methods of biochemical and structural analysis.
The problems of phenotypic analysis of plants
mutant for 14-3-3 proteins. Phenotypic analysis of
mutants could provide opportunities for identification
and characterization of specific functions of 14-3-3 iso-
forms in plants. The knockouts, overexpression, and
heterologous expression of 14-3-3 proteins in differ-
ent plant species have been extensively investigated
[18,  25,  37-45,  169], however, the studies on 14-3-3 mu-
tants are often unsystematic and incomplete. The lack
of systematic approach is evidenced by the fact that
there are no published studies where single mutants
for all 13 isoforms in A. thaliana have been pheno-
typically characterized. The isoforms for studying
are often chosen randomly; an isoform of one plant
can be expressed in another plant. The studies are
often focused on a particular trait of mutant plants
WORLD OF PLANT 14-3-3 PROTEINS S25
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
and ignore the others (there are very few or no stud-
ies of the “one isoform-many traits” or “one trait-all
isoforms” type). As noted above, the main phenotyp-
ic manifestations of 14-3-3 mutations are changes in
plant growth, timing of transition to flowering, toler-
ance to salt stress, and starch accumulation, i.e., traits
important in agriculture. We assume that the authors
of such studies reported only prominent and omitted
less obvious effects of mutations. Interpretation of
results of knockout studies can be significantly com-
plicated by the compensatory effect of other 14-3-3
isoforms, which certainly occurs in the case of fairly
similar 14-3-3 proteins. The scope and fundamental
significance of studies on the reverse genetics of plant
14-3-3 proteins could be greatly increased by using
a systematic approach and comprehensive analysis
of 14-3-3 mutants.
Contributions. I. A. S. and N. N. S. conceived the
idea of critically summarizing current literature and
discussed the plan; I. A. S. analyzed the literature,
wrote the review, and prepared tables and illustra-
tions with input from N. N. S.; I. A. S., and N. N. S. edit-
ed the manuscript.
Funding. The work was in part supported by the
Russian Science Foundation (project no.24-74-00091).
Ethics approval and consent to participate.
This work does not contain any studies involving hu-
man and animal subjects.
Conflict of interest. The authors of this work de-
clare that they have no conflicts of interest.
REFERENCES
1. Cornell,B., and Toyo-oka,K. (2017) 14-3-3 proteins in
brain development: neurogenesis, neuronal migration
and neuromorphogenesis, Front. Mol. Neurosci., 10,
318, https://doi.org/10.3389/fnmol.2017.00318.
2. Moore, B., Perez, V., and Carlson, F. (1967) Physiolog-
ical and biochemical aspects of nervous integration,
Englewood Cliffs NJ Prentice-Hall, pp.343-359.
3. Aitken, A. (2006) 14-3-3 proteins: a historic over-
view, Semin. Cancer Biol., 16, 162-172, https://doi.org/
10.1016/j.semcancer.2006.03.005.
4. Brandt, J., Thordal-Christensen, H., Vad, K.,
Gregersen, P., and Collinge, D. (1992) A pathogen-in-
duced gene of barley encodes a protein showing
high similarity to a protein kinase regulator, Plant J.,
2, 815-820, https://doi.org/10.1046/j.1365-313X.1992.
t01-18-00999.x.
5. De Vetten, N.C., Lu,G., and Feri, R.J. (1992) A maize
protein associated with the G-box binding com-
plex has homology to brain regulatory proteins,
Plant Cell, 4, 1295-1307, https://doi.org/10.1105/tpc.
4.10.1295.
6. Hirsch,S., Aitken,A., Bertsch,U., and Soll,J. (1992) A
plant homologue to mammalian brain 14-3-3 protein
and protein kinase C inhibitor, FEBS Lett., 296, 222-
224, https://doi.org/10.1016/0014-5793(92)80384-S.
7. Lu, G., DeLisle, A. J., De Vetten, N. C., and Ferl, R. J.
(1992) Brain proteins in plants: an Arabidopsis homo-
log to neurotransmitter pathway activators is part of
a DNA binding complex, Proc. Natl. Acad. Sci. USA, 89,
11490-11494, https://doi.org/10.1073/pnas.89.23.11490.
8. Korthout, H.A.A.J., and de Boer, A.H. (1994) A fusi-
coccin binding protein belongs to the family of 14-3-
3 brain protein homologs, Plant Cell, 6, 1681, https://
doi.org/10.2307/3869953.
9. Daugherty, C.J., Rooney, M.F., Miller, P.W., and Ferl,
R.J. (1996) Molecular organization and tissue-specific
expression of an Arabidopsis 14-3-3 gene, Plant Cell,
8, 1239-1248, https://doi.org/10.1105/tpc.8.8.1239.
10. Bachmann, M., Huber, J. L., Liao, P.-C., Gage, D. A.,
and Huber, S.C. (1996) The inhibitor protein of phos-
phorylated nitrate reductase from spinach (Spinacia
oleracea) leaves is a 14-3-3 protein, FEBS Lett., 387,
127-131, https://doi.org/10.1016/0014-5793(96)00478-4.
11. Jia,C., Guo,B., Wang,B., Li,X., Yang,T., Li,N., Wang,J.,
and Yu,Q. (2022) Genome-wide identification and ex-
pression analysis of the 14-3-3 (TFT) gene family in
tomato, and the role of SlTFT4 in salt stress, Plants,
11, 3491, https://doi.org/10.3390/plants11243491.
12. XU, W.F., and SHI, W. M. (2006) Expression profiling
of the 14-3-3 gene family in response to salt stress
and potassium and iron deficiencies in young tomato
(Solanum lycopersicum) roots: analysis by real-time
RT-PCR, Ann. Bot., 98, 965-974, https://doi.org/10.1093/
aob/mcl189.
13. Jarillo, J. A., Capel, J., Leyva, A., Martínez-Zapater,
J.M., and Salinas,J. (1994) Two related low-tempera-
ture-inducible genes of Arabidopsis encode proteins
showing high homology to 14-3-3 proteins, a family
of putative kinase regulators, Plant Mol. Biol., 25,
693-704, https://doi.org/10.1007/BF00029607.
14. Mikhaylova, Y.V., Puzanskiy, R.K., and Shishova, M.F.
(2021) Evolution of 14-3-3 proteins in angiosperm
plants: recurring gene duplication and loss, Plants,
10, 2724, https://doi.org/10.3390/plants10122724.
15. Yao,Y., Du,Y., Jiang,L., and Liu, J.Y. (2007) Molecular
analysis and expression patterns of the 14-3-3 gene
family from Oryza sativa, J. Biochem. Mol. Biol., 40,
349-357, https://doi.org/10.5483/bmbrep.2007.40.3.349.
16. Zhang, Z.-B., Wang, X.-K., Wang,S., Guan,Q., Zhang,W.,
and Feng, Z.-G. (2022) Expansion and diversification
of the 14-3-3 gene family in Camellia sinensis, J.Mol.
Evol., 90, 296-306, https://doi.org/10.1007/s00239-022-
10060-6.
17. Li,X., and Dhaubhadel,S. (2011) Soybean 14-3-3 gene
family: identification and molecular characterization,
Planta, 233, 569-582, https://doi.org/10.1007/s00425-
010-1315-6.
SEDLOV, SLUCHANKOS26
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
18. Ren, J., Zhang, P., Dai, Y., Liu, X., Lu, S., Guo, L.,
Gou, H., and Mao, J. (2023) Evolution of the 14-3-3
gene family in monocotyledons and dicotyledons and
validation of MdGRF13 function in transgenic, Arabi-
dopsis thaliana, Plant Cell Rep., 42, 1345-1364, https://
doi.org/10.1007/s00299-023-03035-4.
19. Cao, H., and Yan, Y. (2016) Molecular characteri-
zation of the 14-3-3 gene family in Brachypodium
distachyonL. Reveals high evolutionary conservation
and diverse responses to abiotic stresses, Front. Plant
Sci., 7, 1099, https://doi.org/10.3389/fpls.2016.01099.
20. Carretero-Paulet,L., and Fares, M.A. (2012) Evolution-
ary dynamics and functional specialization of plant
paralogs formed by whole and small-scale genome
duplications, Mol. Biol. Evol., 29, 3541-3551, https://
doi.org/10.1093/molbev/mss162.
21. Landis, J. B., Soltis, D. E., Li, Z., Marx, H. E., Barker,
M. S., Tank, D. C., and Soltis, P. S. (2018) Impact of
whole-genome duplication events on diversification
rates in angiosperms, Am.J. Bot., 105, 348-363, https://
doi.org/10.1002/ajb2.1060.
22. Zuo, X., Wang, S., Xiang, W., Yang, H., Tahir, M. M.,
Zheng, S., An, N., Han, M., Zhao, C., and Zhang, D.
(2021) Genome-wide identification of the 14-3-3 gene
family and its participation in floral transition by in-
teracting with TFL1/FT in apple, BMC Genomics, 22,
41, https://doi.org/10.1186/s12864-020-07330-2.
23. DeLille, J. M., Sehnke, P. C., and Ferl, R. J. (2001)
The Arabidopsis 14-3-3 family of signaling regula-
tors, Plant Physiol., 126, 35-38, https://doi.org/10.1104/
pp.126.1.35.
24. Yashvardhini,N., Bhattacharya,S., Chaudhuri,S., and
Sengupta, D. N. (2018) RETRACTED Molecular char-
acterization of the 14-3-3 gene family in rice and its
expression studies under abiotic stress, Planta, 247,
229-253, https://doi.org/10.1007/s00425-017-2779-4.
25. Liu,J., Jiang,C., Kang,L., Zhang,H., Song,Y., Zou,Z.,
and Zheng,W. (2020) Over-expression of a 14-3-3 pro-
tein from foxtail millet improves plant tolerance to
salinity stress in Arabidopsis thaliana, Front. Plant
Sci., 11, 449, https://doi.org/10.3389/fpls.2020.00449.
26. Obergfell, E., Hohmann, U., Moretti, A., Chen, H.,
and Hothorn, M. (2024) Mechanistic insights into
the function of 14-3-3 proteins as negative regula-
tors of brassinosteroid signaling in Arabidopsis, Plant
Cell Physiol., 65, 1674-1688, https://doi.org/10.1093/
pcp/pcae056.
27. Ashkenazy, H., Abadi, S., Martz, E., Chay, O.,
Mayrose, I., Pupko, T., and Ben-Tal, N. (2016) Con-
Surf 2016: an improved methodology to estimate
and visualize evolutionary conservation in macro-
molecules, Nucleic Acids Res., 44, W344-W350, https://
doi.org/10.1093/nar/gkw408.
28. The Angiosperm Phylogeny Group (2016) An update
of the Angiosperm Phylogeny Group classification for
the orders and families of flowering plants: APG IV,
Bot. J. Linn. Soc., 181, 1-20, https://doi.org/10.1111/
boj.12385.
29. Van Heusden, G. P. H., Wenzel, T. J., Lagendijk, E. L.,
de Steensma, H. Y., and van den Berg, J. A. (1992)
Characterization of the yeast BMH1 gene encoding
a putative protein homologous to mammalian pro-
tein kinase II activators and protein kinase C inhib-
itors, FEBS Lett., 302, 145-150, https://doi.org/10.1016/
0014-5793(92)80426-H.
30. van Heusden, G. P. H., and Yde Steensma, H. (2006)
Yeast 14-3-3 proteins, Yeast, 23, 159-171, https://
doi.org/10.1002/yea.1338.
31. Van Heusden, G. P. H., Griffiths, D. J. F., Ford, J. C.,
Chin-A-Woeng, T. F. C., Schrader, P. A. T., Carr, A. M.,
and Steensma, H.Y. (1995) The 14-3-3 proteins encod-
ed by the BMH1 and BMH2 genes are essential in the
yeast Saccharomyces cerevisiae and can be replaced
by a plant homologue, Eur. J. Biochem., 229, 45-53,
https://doi.org/10.1111/j.1432-1033.1995.0045l.x.
32. van Heusden, G.P.H., van der Zanden, A.L., Ferl, R.J.,
and Steensma, H.Y. (1996) Four Arabidopsis thaliana
14-3-3 protein isoforms can complement the lethal
yeast bmh1 bmh2 double disruption, FEBS Lett., 391,
252-256, https://doi.org/10.1016/0014-5793(96)00746-6.
33. Kumar, K., Muthamilarasan, M., Bonthala, V. S.,
Roy,R., and Prasad,M. (2015) Unraveling 14-3-3 pro-
teins in C4 panicoids with emphasis on model plant
setaria italica reveals phosphorylation-dependent
subcellular localization of RS splicing factor, PLoS
One, 10, e0123236, https://doi.org/10.1371/journal.
pone.0123236.
34. Tian,F., Wang, T., Xie, Y., Zhang, J., and Hu, J. (2015)
Genome-wide identification, classification, and ex-
pression analysis of 14-3-3 gene family in popu-
lus, PLoS One, 10, e0123225, https://doi.org/10.1371/
journal.pone.0123225.
35. Wang, Y., Xu, Q., Shan, H., Ni, Y., Xu, M., Xu, Y.,
Cheng,B., and Li,X. (2023) Genome-wide analysis of
14-3-3 gene family in four gramineae and its response
to mycorrhizal symbiosis in maize, Front. Plant Sci.,
14, 1117879, https://doi.org/10.3389/fpls.2023.1117879.
36. Van Kleeff, P. J. M., Jaspert, N., Li, K. W., Rauch, S.,
Oecking, C., and de Boer, A. H. (2014) Higher order
Arabidopsis 14-3-3 mutants show 14-3-3 involvement
in primary root growth both under control and abiot-
ic stress conditions, J.Exp. Bot., 65, 5877-5888, https://
doi.org/10.1093/jxb/eru338.
37. Keicher, J., Jaspert, N., Weckermann, K., Möller, C.,
Throm, C., Kintzi, A., and Oecking, C. (2017) Arabi-
dopsis 14-3-3 epsilon members contribute to polari-
ty of PIN auxin carrier and auxin transport-related
development, eLife, 6, e24336, https://doi.org/10.7554/
eLife.24336.
38. Mayfield, J. D., Paul, A.-L., and Ferl, R. J. (2012) The
14-3-3 proteins of Arabidopsis regulate root growth
and chloroplast development as components of the
WORLD OF PLANT 14-3-3 PROTEINS S27
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
photosensory system, J. Exp. Bot., 63, 3061-3070,
https://doi.org/10.1093/jxb/ers022.
39. He,Y., Wu,J., Lv,B., Li,J., Gao,Z., Xu,W., Baluška,F.,
Shi,W., Shaw, P.C., and Zhang,J. (2015) Involvement
of 14-3-3 protein GRF9 in root growth and response
under polyethylene glycol-induced water stress, J.Exp.
Bot., 66, 2271-2281, https://doi.org/10.1093/jxb/erv149.
40. Mayfield, J.D., Folta, K.M., Paul, A.-L., and Ferl, R.J.
(2007) The 14-3-3 proteins μ and υ influence transi-
tion to flowering and early phytochrome response,
Plant Physiol., 145, 1692-1702, https://doi.org/10.1104/
pp.107.108654.
41. Purwestri, Y. A., Ogaki, Y., Tamaki, S., Tsuji, H., and
Shimamoto, K. (2009) The 14-3-3 protein GF14c acts
as a negative regulator of flowering in rice by inter-
acting with the florigen Hd3a, Plant Cell Physiol., 50,
429-438, https://doi.org/10.1093/pcp/pcp012.
42. Wang, X., Chang, L., Tong, Z., Wang, D., Yin, Q.,
Wang,D., Jin,X., Yang,Q., Wang,L., Sun,Y., Huang,Q.,
Guo,A., and Peng,M. (2016) Proteomics profiling re-
veals carbohydrate metabolic enzymes and 14-3-3
Proteins play important roles for starch accumulation
during cassava root tuberization, Sci. Rep., 6, 19643,
https://doi.org/10.1038/srep19643.
43. Pan,R., Wang,Y., An,F., Yao,Y., Xue,J., Zhu,W., Luo,X.,
Lai, H., and Chen, S. (2023) Genome-wide identifica-
tion and characterization of 14-3-3 gene family relat-
ed to negative regulation of starch accumulation in
storage root of Manihot esculenta, Front. Plant Sci., 14,
1184903, https://doi.org/10.3389/fpls.2023.1184903.
44. Sehnke, P.C., Chung, H.-J., Wu,K., and Ferl, R.J. (2001)
Regulation of starch accumulation by granule-associ-
ated plant 14-3-3 proteins, Proc. Natl. Acad. Sci. USA,
98, 765-770, https://doi.org/10.1073/pnas.98.2.765.
45. Zhu,Y., Kuang,W., Leng,J., Wang,X., Qiu,L., Kong,X.,
Wang, Y., and Zhao, Q. (2023) The apple 14-3-3 gene
MdGRF6 negatively regulates salt tolerance, Front.
Plant Sci., 14, 1161539, https://doi.org/10.3389/fpls.
2023.1161539.
46. Paul, A. L., Denison, F. C., Schultz, E. R., Zupanska,
A.K., and Ferl, R.J. (2012) 14-3-3 Phosphoprotein in-
teraction networks – does isoform diversity present
functional interaction specification? Front. Plant Sci.,
3, 190, https://doi.org/10.3389/fpls.2012.00190.
47. Rosenquist, M., Sehnke, P., Ferl, R. J., Sommarin, M.,
and Larsson, C. (2000) Evolution of the 14-3-3 pro-
tein family: does the large number of isoforms in
multicellular organisms reflect functional specifici-
ty? J. Mol. Evol., 51, 446-458, https://doi.org/10.1007/
s002390010107.
48. Rittinger, K., Budman, J., Xu, J., Volinia, S., Cantley,
L. C., Smerdon, S. J., Gamblin, S. J., and Yaffe, M. B.
(1999) Structural analysis of 14-3-3 phosphopeptide
complexes identifies a dual role for the nuclear export
signal of 14-3-3 in ligand binding, Mol. Cell, 4, 153-166,
https://doi.org/10.1016/S1097-2765(00)80363-9.
49. Jiang,W., Tong,T., Li,W., Huang,Z., Chen,G., Zeng,F.,
Riaz, A., Amoanimaa-Dede, H., Pan, R., Zhang, W.,
Deng, F., and Chen, Z.-H. (2022) Molecular Evolution
of plant 14-3-3 proteins and function of Hv14-3-3A in
stomatal regulation and drought tolerance, Plant Cell
Physiol., 63, 1857-1872, https://doi.org/10.1093/pcp/
pcac034.
50. Lopez-Girona, A., Furnari, B., Mondesert, O., and
Russell,P. (1999) Nuclear localization of Cdc25 is reg-
ulated by DNA damage and a 14-3-3 protein, Nature,
397, 172-175, https://doi.org/10.1038/16488.
51. Su, C.-H., Zhao, R., Velazquez-Torres, G., Chen, J.,
Gully, C., Yeung, S.-C. J., and Lee, M.-H. (2010) Nucle-
ar export regulation of COP1 by 14-3-3σ in response
to DNA damage, Mol. Cancer, 9, 243, https://doi.
org/10.1186/1476-4598-9-243.
52. Kumagai,A., and Dunphy, W.G. (1999) Binding of 14-
3-3 proteins and nuclear export control the intracel-
lular localization of the mitotic inducer Cdc25, Genes
Dev., 13, 1067-1072.
53. Sluchanko, N. N., and Gusev, N. B. (2017) Moonlight-
ing chaperone-like activity of the universal regula-
tory 14-3-3 proteins, FEBS J., 284, 1279-1295, https://
doi.org/10.1111/febs.13986.
54. Shen, W., Clark, A. C., and Huber, S. C. (2003) The
C-terminal tail of Arabidopsis 14-3-3ω functions as
an autoinhibitor and may contain a tenth α-helix,
Plant J., 34, 473-484, https://doi.org/10.1046/j.1365-
313X.2003.01739.x.
55. Truong, A. B., Masters, S. C., Yang, H., and Fu, H.
(2002) Role of the 14-3-3 C-terminal loop in ligand in-
teraction, Proteins Struct. Funct. Bioinforma, 49, 321-
325, https://doi.org/10.1002/prot.10210.
56. Silhan,J., Obsilova, V., Vecer, J., Herman, P., Sulc,M.,
Teisinger,J., and Obsil,T. (2004) 14-3-3 Protein C-termi-
nal stretch occupies ligand binding groove and is dis-
placed by phosphopeptide binding, J.Biol. Chem., 279,
49113-49119, https://doi.org/10.1074/jbc.M408671200.
57. Athwal, G.S., and Huber, S.C. (2002) Divalent cations
and polyamines bind to loop 8 of 14-3-3 proteins,
modulating their interaction with phosphorylated ni-
trate reductase, Plant J., 29, 119-129, https://doi.org/
10.1046/j.0960-7412.2001.01200.x.
58. Pallucca, R., Visconti, S., Camoni, L., Cesareni, G.,
Melino,S., Panni,S., Torreri,P., and Aducci,P. (2014)
Specificity of ε and non-ε isoforms of Arabidopsis 14-
3-3 proteins towards the H
+
-ATPase and other targets,
PLoS One, 9, e90764, https://doi.org/10.1371/journal.
pone.0090764.
59. Börnke, F. (2005) The variable C-terminus of 14-3-3
proteins mediates isoform-specific interaction with
sucrose-phosphate synthase in the yeast two-hybrid
system, J. Plant Physiol., 162, 161-168, https://doi.org/
10.1016/j.jplph.2004.09.006.
60. Lu,G., Sehnke, P.C., and Ferl, R.J. (1994) Phosphoryla-
tion and calcium binding properties of an Arabidopsis
SEDLOV, SLUCHANKOS28
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
GF14 brain protein homolog, Plant Cell, 6, 501,
https://doi.org/10.2307/3869930.
61. Athwal, G. S., Huber, J. L., and Huber, S.C. (1998) Bi-
ological significance of divalent metal ion binding to
14-3-3 proteins in relationship to nitrate reductase
inactivation, Plant Cell Physiol., 39, 1065-1072, https://
doi.org/10.1093/oxfordjournals.pcp.a029303.
62. Yang,Z., Wang,C., Xue,Y., Liu,X., Chen,S., Song, C.P.,
Yang, Y., and Guo, Y. (2019) Calcium-activated 14-3-3
proteins as a molecular switch in salt stress toler-
ance, Nat. Commun., 10, 1199, https://doi.org/10.1038/
s41467-019-09181-2.
63. Day, I.S., Reddy, V.S., Shad Ali,G., and Reddy, A.S.N.
(2002) Analysis of EF-hand-containing proteins in
Arabidopsis, Genome Biol., 3, RESEARCH0056, https://
doi.org/10.1186/gb-2002-3-10-research0056.
64. Manak, M. S., and Ferl, R. J. (2007) Divalent cation
effects on interactions between multiple Arabidopsis
14-3-3 isoforms and phosphopeptide targets, Bio-
chemistry, 46, 1055-1063, https://doi.org/10.1021/
bi061366c.
65. Trošanová, Z., Louša, P., Kozeleková, A., Brom, T.,
Gašparik,N., Tungli,J., Weisová,V., Župa,E., Žoldák,G.,
and Hritz, J. (2022) Quantitation of human 14-3-3 ζ
dimerization and the effect of phosphorylation on
dimer-monomer equilibria, J.Mol. Biol., 434, 167479,
https://doi.org/10.1016/j.jmb.2022.167479.
66. Chaudhri, M., Scarabel, M., and Aitken, A. (2003)
Mammalian and yeast 14-3-3 isoforms form distinct
patterns of dimers invivo, Biochem. Biophys. Res.
Commun., 300, 679-685, https://doi.org/10.1016/S0006-
291X(02)02902-9.
67. Yang, X., Lee, W. H., Sobott, F., Papagrigoriou, E.,
Robinson, C. V., Grossmann, J. G., Sundström, M.,
Doyle, D. A., and Elkins, J.M. (2006) Structural basis
for protein-protein interactions in the 14-3-3 protein
family, Proc. Natl. Acad. Sci. USA, 103, 17237-17242,
https://doi.org/10.1073/pnas.0605779103.
68. Verdoodt, B., Benzinger, A., Popowicz, G. M., Holak,
T. A., and Hermeking, H. (2006) Characterization of
14-3-3sigma dimerization determinants: requirement
of homodimerization for inhibition of cell prolifera-
tion, Cell Cycle, 5, 2920-2926, https://doi.org/10.4161/
cc.5.24.3571.
69. Wilker, E. W., Grant, R. A., Artim, S. C., and Yaffe,
M. B. (2005) A structural basis for 14-3-3σ function-
al specificity, J. Biol. Chem., 280, 18891-18898, https://
doi.org/10.1074/jbc.M500982200.
70. Torosyan,H., Paul, M.D., Forget,A., Lo,M., Diwanji,D.,
Pawłowski,K., Krogan, N.J., Jura,N., and Verba, K.A.
(2023) Structural insights into regulation of the PEAK3
pseudokinase scaffold by 14-3-3, Nat. Commun., 14,
3543, https://doi.org/10.1038/s41467-023-38864-0.
71. Tugaeva, K. V., Titterington, J., Sotnikov, D. V.,
Maksimov, E. G., Antson, A. A., and Sluchanko,
N. N. (2020) Molecular basis for the recognition of
steroidogenic acute regulatory protein by the 14-
3-3 protein family, FEBS J., 287, 3944-3966, https://
doi.org/10.1111/febs.15474.
72. Ghorbani, S., Fossbakk,A., Jorge-Finnigan,A., Flydal,
M. I., Haavik, J., and Kleppe, R. (2016) Regulation of
tyrosine hydroxylase is preserved across different
homo- and heterodimeric 14-3-3 proteins, Amino
Acids, 48, 1221-1229, https://doi.org/10.1007/s00726-
015-2157-0.
73. Wu, K., Lu, G., Sehnke, P., and Ferl, R. J. (1997) The
heterologous interactions among plant 14-3-3 pro-
teins and identification of regions that are important
for dimerization, Arch. Biochem. Biophys., 339, 2-8,
https://doi.org/10.1006/abbi.1996.9841.
74. Gökirmak, T., Denison, F. C., Laughner, B. J., Paul,
A.L., and Ferl, R.J. (2015) Phosphomimetic mutation
of a conserved serine residue in Arabidopsis thali-
ana 14-3-3ω suggests a regulatory role of phosphor-
ylation in dimerization and target interactions, Plant
Physiol. Biochem., 97, 296-303, https://doi.org/10.1016/
j.plaphy.2015.10.022.
75. Oh, C.-S., and Martin, G.B. (2011) Tomato 14-3-3 pro-
tein TFT7 interacts with a MAP kinase kinase to regu-
late immunity-associated programmed cell death me-
diated by diverse disease resistance proteins, J. Biol.
Chem., 286, 14129-14136, https://doi.org/10.1074/
jbc.M111.225086.
76. Paul, A.-L., Sehnke, P. C., and Ferl, R. J. (2005) Iso-
form-specific subcellular localization among 14-3-3
proteins in Arabidopsis seems to be driven by client
interactions, Mol. Biol. Cell, 16, 1735-1743, https://
doi.org/10.1091/mbc.e04-09-0839.
77. Zhang, Z.-T., Zhou, Y., Li, Y., Shao, S.-Q., Li, B.-Y., Shi,
H.-Y., and Li, X.-B. (2010) Interactome analysis of the
six cotton 14-3-3s that are preferentially expressed in
fibres and involved in cell elongation, J.Exp. Bot., 61,
3331-3344, https://doi.org/10.1093/jxb/erq155.
78. Woodcock, J. M., Murphy, J., Stomski, F. C., Berndt,
M. C., and Lopez, A. F. (2003) The dimeric versus
monomeric status of 14-3-3ζ is controlled by phos-
phorylation of Ser58 at the dimer interface, J. Biol.
Chem., 278, 36323-36327, https://doi.org/10.1074/
jbc.M304689200.
79. Denison, F.C., Gökirmak,T., and Ferl, R.J. (2014) Phos-
phorylation-related modification at the dimer inter-
face of 14-3-3ω dramatically alters monomer inter-
action dynamics, Arch. Biochem. Biophys., 541, 1-12,
https://doi.org/10.1016/j.abb.2013.10.025.
80. Sluchanko, N. N., Chernik, I. S., Seit-Nebi, A. S.,
Pivovarova, A. V., Levitsky, D. I., and Gusev, N. B.
(2008) Effect of mutations mimicking phosphory-
lation on the structure and properties of human
14-3-3ζ, Arch. Biochem. Biophys., 477, 305-312, https://
doi.org/10.1016/j.abb.2008.05.020.
81. Swatek, K. N., Wilson, R. S., Ahsan, N., Tritz, R. L.,
and Thelen, J. J. (2014) Multisite phosphorylation of
WORLD OF PLANT 14-3-3 PROTEINS S29
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
14-3-3 proteins by calcium-dependent protein kinas-
es, Biochem. J., 459, 15-25, https://doi.org/10.1042/
BJ20130035.
82. Lachaud, C., Prigent, E., Thuleau, P., Grat, S., Da
Silva, D., Brière, C., Mazars,C., and Cotelle, V. (2013)
14-3-3-Regulated Ca
2+
-dependent protein kinase CPK3
is required for sphingolipid-induced cell death in
Arabidopsis, Cell Death Differ., 20, 209-217, https://
doi.org/10.1038/cdd.2012.114.
83. de Boer, A. H., van Kleeff, P.J. M., and Gao,J. (2013)
Plant 14-3-3 proteins as spiders in a web of phosphor-
ylation, Protoplasma, 250, 425-440, https://doi.org/
10.1007/s00709-012-0437-z.
84. Bustos, D. M., and Iglesias, A. A. (2006) Intrinsic dis-
order is a key characteristic in partners that bind
14-3-3 proteins, Proteins Struct. Funct. Bioinforma, 63,
35-42, https://doi.org/10.1002/prot.20888.
85. Karlberg, T., Hornyak, P., Pinto, A. F., Milanova, S.,
Ebrahimi,M., Lindberg,M., Püllen,N., Nordström,A.,
Löverli, E., Caraballo, R., Wong, E. V., Näreoja, K.,
Thorsell, A.-G., Elofsson, M., De La Cruz, E. M.,
Björkegren,C., and Schüler,H. (2018) 14-3-3 proteins
activate Pseudomonas exotoxins-S and -T by chaper-
oning a hydrophobic surface, Nat. Commun., 9, 3785,
https://doi.org/10.1038/s41467-018-06194-1.
86. Coblitz,B., Wu,M., Shikano,S., and Li,M. (2006) C-ter-
minal binding: An expanded repertoire and function
of 14-3-3 proteins, FEBS Lett., 580, 1531-1535, https://
doi.org/10.1016/j.febslet.2006.02.014.
87. Johnson, C., Crowther, S., Stafford, M. J., Campbell,
D. G., Toth, R., and MacKintosh, C. (2010) Bioinfor-
matic and experimental survey of 14-3-3-binding
sites, Biochem. J., 427, 69-78, https://doi.org/10.1042/
BJ20091834.
88. Würtele,M., Jelich-Ottmann,C., Wittinghofer,A., and
Oecking, C. (2003) Structural view of a fungal toxin
acting on a 14-3-3 regulatory complex, EMBO J., 22,
987-994, https://doi.org/10.1093/emboj/cdg104.
89. Stevers, L.M., Lam, C.V., Leysen, S.F.R., Meijer, F.A.,
van Scheppingen, D.S., de Vries, R. M. J. M., Carlile,
G.W., Milroy, L.G., Thomas, D.Y., Brunsveld, L., and
Ottmann, C. (2016) Characterization and small-mol-
ecule stabilization of the multisite tandem bind-
ing between 14-3-3 and the R domain of CFTR,
Proc. Natl. Acad. Sci. USA, 113, E1152-E1161, https://
doi.org/10.1073/pnas.1516631113.
90. Centorrino,F., Ballone,A., Wolter,M., and Ottmann,C.
(2018) Biophysical and structural insight into the
USP8/14-3-3 interaction, FEBS Lett., 592, 1211-1220,
https://doi.org/10.1002/1873-3468.13017.
91. Manschwetus, J. T., Wallbott, M., Fachinger, A.,
Obergruber,C., Pautz,S., Bertinetti,D., Schmidt, S.H.,
and Herberg, F. W. (2020) Binding of the human
14-3-3 isoforms to distinct sites in the leucine-rich
repeat kinase 2, Front. Neurosci., 14, 302, https://
doi.org/10.3389/fnins.2020.00302.
92. Gogl, G., Tugaeva, K. V., Eberling, P., Kostmann, C.,
Trave, G., and Sluchanko, N. N. (2021) Hierarchized
phosphotarget binding by the seven human 14-3-3 iso-
forms, Nat. Commun., 12, 2-13, https://doi.org/10.1038/
s41467-021-21908-8.
93. Bachmann,M., Huber, J.L., Athwal, G.S., Wu,K., Ferl,
R.J., and Huber, S.C. (1996) 14-3-3 proteins associate
with the regulatory phosphorylation site of spinach
leaf nitrate reductase in an isoform-specific manner
and reduce dephosphorylation of Ser-543 by endog-
enous protein phosphatases, FEBS Lett., 398, 26-30,
https://doi.org/10.1016/S0014-5793(96)01188-X.
94. Lambeck, I., Chi, J. C., Krizowski, S., Mueller, S.,
Mehlmer, N., Teige, M., Fischer, K., and Schwarz, G.
(2010) Kinetic analysis of 14-3-3-inhibited arabidop-
sis thaliana nitrate reductase, Biochemistry, 49, 8177-
8186, https://doi.org/10.1021/bi1003487.
95. Nishiyama, K., Aihara, Y., Suzuki, T., Takahashi, K.,
Kinoshita, T., Dohmae, N., Sato, A., and Hagihara, S.
(2024) Discovery of a plant 14-3-3 inhibitor possess-
ing isoform selectivity and in planta activity, Angew.
Chem. Int. Ed., 63, e202400218, https://doi.org/10.1002/
anie.202400218.
96. Taoka, K.I., Ohki,I., Tsuji,H., Furuita,K., Hayashi,K.,
Yanase, T., Yamaguchi, M., Nakashima, C., Purwestri,
Y.A., Tamaki,S., Ogaki,Y., Shimada,C., Nakagawa,A.,
Kojima, C., and Shimamoto, K. (2011) 14-3-3 pro-
teins act as intracellular receptors for rice Hd3a fl-
origen, Nature, 476, 332-335, https://doi.org/10.1038/
nature10272.
97. Lambeck, I.C., Fischer-Schrader,K., Niks,D., Roeper,J.,
Chi, J.-C., Hille, R., and Schwarz, G. (2012) Molecular
mechanism of 14-3-3 protein-mediated inhibition of
plant nitrate reductase, J.Biol. Chem., 287, 4562-4571,
https://doi.org/10.1074/jbc.M111.323113.
98. Toroser, D., Athwal, G. S., and Huber, S. C. (1998)
Site-specific regulatory interaction between spinach
leaf sucrose-phosphate synthase and 14-3-3 proteins,
FEBS Lett., 435, 110-114, https://doi.org/10.1016/S0014-
5793(98)01048-5.
99. Ottmann, C., Marco, S., Jaspert, N., Marcon, C.,
Schauer,N., Weyand,M., Vandermeeren,C., Duby,G.,
Boutry, M., Wittinghofer, A., Rigaud, J. L., and
Oecking, C. (2007) Structure of a 14-3-3 coordinated
hexamer of the plant plasma membrane H
+
-ATPase
by combining X-ray crystallography and electron cry-
omicroscopy, Mol. Cell, 25, 427-440, https://doi.org/
10.1016/j.molcel.2006.12.017.
100. Finnemann,J., and Schjoerring, J.K. (2000) Post-trans-
lational regulation of cytosolic glutamine synthetase
by reversible phosphorylation and 14-3-3 protein in-
teraction, PlantJ., 24, 171-181, https://doi.org/10.1046/
j.1365-313X.2000.00863.x.
101. Ryu, H., Cho, H., Kim,K., and Hwang, I. (2010) Phos-
phorylation dependent nucleocytoplasmic shuttling
of BES1 is a key regulatory event in brassinosteroid
SEDLOV, SLUCHANKOS30
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
signaling, Mol. Cells, 29, 283-290, https://doi.org/
10.1007/s10059-010-0035-x.
102. Gampala, S.S., Kim, T.W., He, J.X., Tang,W., Deng,Z.,
Bai, M. Y., Guan, S., Lalonde, S., Sun, Y., Gendron,
J.M., Chen,H., Shibagaki,N., Ferl, R.J., Ehrhardt, D.,
Chong, K., Burlingame, A. L., and Wang, Z. Y. (2007)
An essential role for 14-3-3 proteins in brassinosteroid
signal transduction in Arabidopsis, Dev. Cell, 13, 177-
189, https://doi.org/10.1016/j.devcel.2007.06.009.
103. Ito, T., Nakata, M., Fukazawa, J., Ishida, S., and
Takahashi, Y. (2014) Scaffold function of Ca
2+
-depen-
dent protein kinase: tobacco Ca
2+
-dependent protein
kinase1 transfers 14-3-3 to the substrate repres-
sion of shoot growth after phosphorylation, Plant
Physiol., 165, 1737-1750, https://doi.org/10.1104/pp.
114.236448.
104. Huang, X., Zhang, Q., Jiang, Y., Yang, C., Wang, Q.,
and Li, L. (2018) Shade-induced nuclear localization
of PIF7 is regulated by phosphorylation and 14-3-
3 proteins in Arabidopsis, eLife, 7, e31636, https://
doi.org/10.7554/eLife.31636.
105. May, T., and Soll, J. (2000) 14-3-3 proteins form
a guidance complex with chloroplast precursor
proteins in plants, Plant Cell, 12, 53, https://doi.
org/10.2307/3871029.
106. Sehnke, P.C., Henry,R., Cline,K., and Ferl, R.J. (2000)
Interaction of a plant 14-3-3 protein with the signal
peptide of a thylakoid-targeted chloroplast precur-
sor protein and the presence of 14-3-3 isoforms in
the chloroplast stroma, Plant Physiol., 122, 235-241,
https://doi.org/10.1104/pp.122.1.235.
107. Fellerer, C., Schweiger, R., Schöngruber, K., Soll, J.,
and Schwenkert,S. (2011) Cytosolic HSP90 cochaper-
ones HOP and FKBP interact with freshly synthesized
chloroplast preproteins of Arabidopsis, Mol. Plant, 4,
1133-1145, https://doi.org/10.1093/mp/ssr037.
108. Law, Y.-S., Ngan, L., Yan, J., Kwok, L. Y., Sun, Y.,
Cheng, S., Schwenkert, S., and Lim, B. L. (2018)
Multiple kinases can phosphorylate the N-terminal
sequences of mitochondrial proteins in Arabidop-
sis thaliana, Front. Plant Sci., 9, 982, https://doi.org/
10.3389/fpls.2018.00982.
109. Law, Y. S., Zhang, R., Guan, X., Cheng, S., Sun, F.,
Duncan,O., Murcha, M.W., Whelan,J., and Lim, B.L.
(2015) Phosphorylation and dephosphorylation of the
presequence of precursor multiple organellar RNA ed-
iting factor 3 during import into mitochondria from
arabidopsis, Plant Physiol, 169, 1344-1355, https://
doi.org/10.1104/pp.15.01115.
110. Schultz, T. F., Medina, J., Hill,A., and Quatrano, R.S.
(1998) 14-3-3 proteins are part of an abscisic acid-
VIVIPAROUS1 (VP1) response complex in the Em pro-
moter and interact with VP1 and EmBP1, Plant Cell,
10, 837-847, https://doi.org/10.1105/tpc.10.5.837.
111. Dong,X., Feng,F., Li,Y., Li, L., Chen, S., and Zhou, J.-
M. (2023) 14-3-3 proteins facilitate the activation of
MAP kinase cascades by upstream immunity-relat-
ed kinases, Plant Cell, 35, 2413-2428, https://doi.org/
10.1093/plcell/koad088.
112. Song,P., Yang,Z., Guo,C., Han,R., Wang,H., Dong,J.,
Kang, D., Guo, Y., Yang, S., and Li, J. (2023) 14-3-3
proteins regulate photomorphogenesis by facilitating
light-induced degradation of PIF3, New Phytol., 237,
140-159, https://doi.org/10.1111/nph.18494.
113. Yoon, G. M., and Kieber, J. J. (2013) 14-3-3 regulates
1-aminocyclopropane-1-carboxylate synthase protein
turnover in Arabidopsis, Plant Cell, 25, 1016-1028,
https://doi.org/10.1105/tpc.113.110106.
114. Ma, W., Kong, Q., Mantyla, J. J., Yang, Y., Ohlrogge,
J. B., and Benning, C. (2016) 14-3-3 protein mediates
plant seed oil biosynthesis through interaction with
AtWRI1, Plant J., 88, 228-235, https://doi.org/10.1111/
tpj.13244.
115. Hausser,A., Link,G., Hoene,M., Russo,C., Selchow,O.,
and Pfizenmaier, K. (2006) Phospho-specific binding
of 14-3-3 proteins to phosphatidylinositol 4-kinase
III β protects from dephosphorylation and stabiliz-
es lipid kinase activity, J. Cell Sci., 119, 3613-3621,
https://doi.org/10.1242/jcs.03104.
116. Dobson, M., Ramakrishnan, G., Ma, S., Kaplun, L.,
Balan,V., Fridman, R., and Tzivion,G. (2011) Bimod-
al regulation of FoxO3 by AKT and 14-3-3, Biochim.
Biophys. Acta Mol. Cell Res., 1813, 1453-1464, https://
doi.org/10.1016/j.bbamcr.2011.05.001.
117. Sluchanko, N. N., Roman, S. G., Chebotareva, N. A.,
and Gusev, N. B. (2014) Chaperone-like activity of
monomeric human 14-3-3ζ on different protein sub-
strates, Arch. Biochem. Biophys., 549, 32-39, https://
doi.org/10.1016/j.abb.2014.03.008.
118. Yano, M., Nakamuta, S., Wu, X., Okumura, Y., and
Kido,H. (2006) A novel function of 14-3-3 protein: 14-
3-3ζ is a heat-shock-related molecular chaperone that
dissolves thermal-aggregated proteins, Mol. Biol. Cell,
17, 4769-4779, https://doi.org/10.1091/mbc.e06-03-0229.
119. Liu, Y.-Q., Liang, C.-Q., Chen, Z.-W., Hu, J., Hu, J.-J.,
Luo, Y.-Y., Chen, Y.-X., and Li, Y.-M. (2023) 14-3-3ζ
participates in the phase separation of phosphory-
lated and glycated tau and modulates the physio-
logical and pathological functions of tau, ACS Chem.
Neurosci., 14, 1220-1225, https://doi.org/10.1021/
acschemneuro.3c00034.
120. Stitt, M., Müller, C., Matt, P., Gibon, Y., Carillo, P.,
Morcuende, R., Scheible, W., and Krapp, A. (2002)
Steps towards an integrated view of nitrogen metab-
olism, J.Exp. Bot., 53, 959-970, https://doi.org/10.1093/
jexbot/53.370.959.
121. MacKintosh,C., and Meek, S.E.M. (2001) Regulation
of plant NR activity by reversible phosphorylation,
14-3-3 proteins and proteolysis, Cell. Mol. Life Sci., 58,
205-214, https://doi.org/10.1007/PL00000848.
122. Lu,G., Campbell, W.H., Schneider,G., and Lindqvist,Y.
(1994) Crystal structure of the FAD-containing
WORLD OF PLANT 14-3-3 PROTEINS S31
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
fragment of corn nitrate reductase at 2.5 Å reso-
lution: relationship to other flavoprotein reduc-
tases, Structure, 2, 809-821, https://doi.org/10.1016/
S0969-2126(94)00082-4.
123. Sinnige, M. P., Roobeek, I., Bunney, T. D., Visser,
A. J. W. G., Mol, J. N. M., and De Boer, A. H. (2005)
Single amino acid variation in barley 14-3-3 proteins
leads to functional isoform specificity in the reg-
ulation of nitrate reductase, Plant J., 44, 1001-1009,
https://doi.org/10.1111/j.1365-313X.2005.02599.x.
124. Chi, J.C., Roeper,J., Schwarz,G., and Fischer- Schrader,K.
(2015) Dual binding of 14-3-3 protein regulates Arabi-
dopsis nitrate reductase activity, J.Biol. Inorg. Chem.,
20, 277-286, https://doi.org/10.1007/s00775-014-1232-4.
125. Sugden, C., Donaghy, P. G., Halford, N. G., and Har-
die, D. G. (1999) Two SNF1-related protein kinases
from spinach leaf phosphorylate and inactivate 3-hy-
droxy-3-methylglutaryl-coenzyme a reductase, nitrate
reductase, and sucrose phosphate synthase invitro,
Plant Physiol., 120, 257-274, https://doi.org/10.1104/
pp.120.1.257.
126. Crepin,N., and Rolland,F. (2019) SnRK1 activation, sig-
naling, and networking for energy homeostasis, Curr.
Opin. Plant Biol., 51, 29-36, https://doi.org/10.1016/j.
pbi.2019.03.006.
127. Broeckx, T., Hulsmans, S., and Rolland, F. (2016) The
plant energy sensor: evolutionary conservation and
divergence of SnRK1 structure, regulation, and func-
tion, J.Exp. Bot., 67, 6215-6252, https://doi.org/10.1093/
jxb/erw416.
128. Pozuelo, M., MacKintosh, C., Galván, A., and
Fernández, E. (2001) Cytosolic glutamine synthe-
tase and not nitrate reductase from the green alga
Chlamydomonas reinhardtii is phosphorylated and
binds 14-3-3 proteins, Planta, 212, 264-269, https://
doi.org/10.1007/s004250000388.
129. Riedel,J., Tischner, R., and Mäck,G. (2001) The chlo-
roplastic glutamine synthetase (GS-2) of tobacco is
phosphorylated and associated with 14-3-3 proteins
inside the chloroplast, Planta, 213, 396-401, https://doi.
org/10.1007/s004250000509.
130. Swatek, K.N., Graham,K., Agrawal, G.K., and Thelen,
J. J. (2011) The 14-3-3 isoforms chi and epsilon dif-
ferentially bind client proteins from developing
Arabidopsis seed, J. Proteome Res., 10, 4076-4087,
https://doi.org/10.1021/pr200263m.
131. Gao, J., van Kleeff, P. J. M., Li, K. W., and de Boer,
A. H. (2021) Physiological and interactomic analy-
sis reveals versatile functions of Arabidopsis 14-3-3
quadruple mutants in response to Fe deficiency,
Sci. Rep., 11, 15551, https://doi.org/10.1038/s41598-
021-94908-9.
132. Paul, A.L., Liu,L., McClung,S., Laughner,B., Chen,S.,
and Ferl, R. J. (2009) Comparative interactomics:
analysis of Arabidopsis 14-3-3 complexes reveals
highly conserved 14-3-3 interactions between humans
and plants, J. Proteome Res., 8, 1913-1924, https://
doi.org/10.1021/pr8008644.
133. Chang, I.-F., Curran, A., Woolsey, R., Quilici, D.,
Cushman, J. C., Mittler, R., Harmon, A., and Harper,
J.F. (2009) Proteomic profiling of tandem affinity pu-
rified 14-3-3 protein complexes in Arabidopsis thali-
ana, Proteomics, 9, 2967-2985, https://doi.org/10.1002/
pmic.200800445.
134. McMichael, R. W., Klein, R. R., Salvucci, M. E., and
Huber, S.C. (1993) Identification of the major regula-
tory phosphorylation site in sucrose-phosphate syn-
thase, Arch. Biochem. Biophys., 307, 248-252, https://
doi.org/10.1006/abbi.1993.1586.
135. Huber, S.C. (2007) Exploring the role of protein phos-
phorylation in plants: from signalling to metabolism,
Biochem. Soc. Trans., 35, 28-32, https://doi.org/10.1042/
BST0350028.
136. Moorhead, G., Douglas, P., Cotelle, V., Harthill, J.,
Morrice,N., Meek,S., Deiting,U., Stitt,M., Scarabel,M.,
Aitken, A., and MacKintosh, C. (1999) Phosphory-
lation-dependent interactions between enzymes of
plant metabolism and 14-3-3 proteins, Plant J., 18,
1-12, https://doi.org/10.1046/j.1365-313X.1999.00417.x.
137. Lutfiyya, L. L., Xu, N., D’Ordine, R. L., Morrell, J. A.,
Miller, P. W., and Duff, S. M. G. (2007) Phylogenet-
ic and expression analysis of sucrose phosphate
synthase isozymes in plants, J. Plant Physiol., 164,
923-933, https://doi.org/10.1016/j.jplph.2006.04.014.
138. Cotelle, V., Meek, S. E. M., Provan, F., Milne, F. C.,
Morrice, N., and MacKintosh, C. (2000) 14-3-3s regu-
late global cleavage of their diverse binding partners
in sugar-starved Arabidopsis cells, EMBOJ., 19, 2869-
2876, https://doi.org/10.1093/emboj/19.12.2869.
139. Gao, J., Van Kleeff, P. J. M., Oecking, C., Li, K. W.,
Erban,A., Kopka,J., Hincha, D.K., and De Boer, A.H.
(2014) Light modulated activity of root alkaline/neu-
tral invertase involves the interaction with 14-3-3
proteins, Plant J., 80, 785-796, https://doi.org/10.1111/
tpj.12677.
140. Bustos, D.M., and Iglesias, A.A. (2003) Phosphorylat-
ed non-phosphorylating glyceraldehyde-3-phosphate
dehydrogenase from heterotrophic cells of wheat
interacts with 14-3-3 proteins, Plant Physiol., 133,
2081-2088, https://doi.org/10.1104/pp.103.030981.
141. Habenicht, A., Quesada, A., and Cerff, R. (1997) Se-
quence of the non-phosphorylating glyceralde-
hyde-3-phosphate dehydrogenase from Nicotiana
plumbaginifolia and phylogenetic origin of the gene
family, Gene, 198, 237-243, https://doi.org/10.1016/
S0378-1119(97)00320-X.
142. Huber, S. C., MacKintosh, C., and Kaiser, W. M.
(2002) Metabolic enzymes as targets for 14-3-3 pro-
teins, Plant Mol. Biol., 50, 1053-1063, https://doi.org/
10.1023/A:1021284002779.
143. Harthill, J. E., Meek, S. E. M., Morrice, N., Peggie,
M. W., Borch, J., Wong, B. H. C., and MacKintosh, C.
SEDLOV, SLUCHANKOS32
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
(2006) Phosphorylation and 14-3-3 binding of Arabi-
dopsis trehalose-phosphate synthase 5 in response
to 2-deoxyglucose, Plant J., 47, 211-223, https://
doi.org/10.1111/j.1365-313X.2006.02780.x.
144. Truesdale, M. R., Toldi, O., and Scott, P. (1999) The
Effect of elevated concentrations of fructose 2,6-bi-
sphosphate on carbon metabolism during deacid-
ification in the crassulacean acid metabolism plant
Kalanchöe daigremontiana1, Plant Physiol., 121, 957-
964, https://doi.org/10.1104/pp.121.3.957.
145. Kulma,A., Villadsen,D., Campbell, D.G., Meek, S.E.M.,
Harthill, J. E., Nielsen, T. H., and MacKintosh, C.
(2004) Phosphorylation and 14-3-3 binding of Arabi-
dopsis 6-phosphofructo-2-kinase/fructose-2,6-bisphos-
phatase, Plant J., 37, 654-667, https://doi.org/10.1111/
j.1365-313X.2003.01992.x.
146. Shin,R., Jez, J.M., Basra,A., Zhang,B., and Schachtman,
D. P. (2011) 14-3-3 proteins fine-tune plant nutrient
metabolism, FEBS Lett., 585, 143-147, https://doi.org/
10.1016/j.febslet.2010.11.025.
147. Cotelle,V., and Leonhardt,N. (2016) 14-3-3 proteins in
guard cell signaling, Front. Plant Sci., 6, 1210, https://
doi.org/10.3389/fpls.2015.01210.
148. Li, Y., Zeng, H., Xu, F., Yan, F., and Xu, W. (2022)
H
+
-ATPases in plant growth and stress responses,
Annu. Rev. Plant Biol., 73, 495-521, https://doi.org/
10.1146/annurev-arplant-102820-114551.
149. Palmgren, M. G., Sommarin, M., Serrano, R., and
Larsson,C. (1991) Identification of an autoinhibitory
domain in the C-terminal region of the plant plasma
membrane H(+)-ATPase, J. Biol. Chem., 266, 20470-
20475, https://doi.org/10.1016/S0021-9258(18)54948-6.
150. Jaspert, N., and Oecking, C. (2002) Regulatory 14-3-3
proteins bind the atypical motif within the C terminus
of the plant plasma membrane H
+
-ATPase via their
typical amphipathic groove, Planta, 216, 136-139,
https://doi.org/10.1007/s00425-002-0915-1.
151. Svennelid,F., Olsson,A., Piotrowski,M., Rosenquist,M.,
Ottman,C., Larsson,C., Oecking,C., and Sommarin,M.
(1999) Phosphorylation of Thr-948 at the C terminus
of the plasma membrane H
+
-ATPase creates a binding
site for the regulatory 14-3-3 protein, Plant Cell, 11,
2379-2391, https://doi.org/10.1105/tpc.11.12.2379.
152. Jelich-Ottmann, C., Weiler, E. W., and Oecking, C.
(2001) Binding of regulatory 14-3-3 proteins to the
Cterminus of the plant plasma membrane H
+
-ATPase
involves part of its autoinhibitory region, J. Biol.
Chem., 276, 39852-39857, https://doi.org/10.1074/
jbc.M106746200.
153. Takahashi, K., Hayashi, K., and Kinoshita, T. (2012)
Auxin activates the plasma membrane H
+
-ATPase
by phosphorylation during hypocotyl elongation
in Arabidopsis, Plant Physiol., 159, 632-641, https://
doi.org/10.1104/pp.112.196428.
154. Pei,D., Hua,D., Deng,J., Wang,Z., Song,C., Wang,Y.,
Wang, Y., Qi, J., Kollist, H., Yang, S., Guo, Y., and
Gong,Z. (2022) Phosphorylation of the plasma mem-
brane H
+
-ATPase AHA2 by BAK1 is required for ABA-
induced stomatal closure in Arabidopsis, Plant Cell,
34, 2708-2729, https://doi.org/10.1093/plcell/koac106.
155. Pertl-Obermeyer,H., Gimeno,A., Kuchler,V., Servili,E.,
Huang, S., Fang, H., Lang, V., Sydow, K., Pöckl, M.,
Schulze, W. X., and Obermeyer, G. (2022) pH mod-
ulates interaction of 14-3-3 proteins with pollen
plasma membrane H
+
ATPases independently from
phosphorylation, J. Exp. Bot., 73, 168-181, https://
doi.org/10.1093/jxb/erab387.
156. Fuglsang, A. T., Visconti, S., Drumm, K., Jahn, T.,
Stensballe, A., Mattei, B., Jensen, O. N., Aducci, P.,
and Palmgren, M. G. (1999) Binding of 14-3-3 pro-
tein to the plasma membrane H
+
-ATPase AHA2 in-
volves the three C-terminal residues Tyr946-Thr-
Val and requires phosphorylation of Thr947, J. Biol.
Chem., 274, 36774-36780, https://doi.org/10.1074/jbc.
274.51.36774.
157. Ottmann,C., Weyand,M., Sassa,T., Inoue,T., Kato,N.,
Wittinghofer, A., and Oecking, C. (2009) A structural
rationale for selective stabilization of anti-tumor inter-
actions of 14-3-3 proteins by cotylenin A, J.Mol. Biol.,
386, 913-919, https://doi.org/10.1016/j.jmb.2009.01.005.
158. Rose, R., Erdmann, S., Bovens, S., Wolf, A., Rose, M.,
Hennig, S., Waldmann, H., and Ottmann, C. (2010)
Identification and structure of small-molecule stabi-
lizers of 14-3-3 protein-protein interactions, Angew.
Chem. Int. Ed., 49, 4129-4132, https://doi.org/10.1002/
anie.200907203.
159. Richter, A., Rose, R., Hedberg, C., Waldmann, H., and
Ottmann, C. (2012) An optimised small-molecule sta-
biliser of the 14-3-3-PMA2 protein-protein interaction,
Chem. Eur. J., 18, 6520-6527, https://doi.org/10.1002/
chem.201103761.
160. Camoni, L., Visconti, S., Aducci, P., and Marra, M.
(2019) From plant physiology to pharmacology: fusi-
coccin leaves the leaves, Planta, 249, 49-57, https://doi.
org/10.1007/s00425-018-3051-2.
161. De Boer, A. H., Watson, B. A., and Cleland, R. E.
(1989) Purification and identification of the fusicoc-
cin binding protein from oat root plasma membrane,
Plant Physiol., 89, 250-259, https://doi.org/10.1104/
pp.89.1.250.
162. Marra, M., Fullone, M. R., Fogliano, V., Pen, J.,
Mattei,M., Masi,S., and Aducci,P. (1994) The 30-kilo-
dalton protein present in purified fusicoccin receptor
preparations is a 14-3-3-like protein, Plant Physiol.,
106, 1497-1501, https://doi.org/10.1104/pp.106.4.1497.
163. Oecking, C., Eckerskorn, C., and Weiler, E. W. (1994)
The fusicoccin receptor of plants is a member of the
14-3-3 superfamily of eukaryotic regulatory proteins,
FEBS Lett., 352, 163-166, https://doi.org/10.1016/0014-
5793(94)00949-X.
164. Paiardini, A., Aducci, P., Cervoni, L., Cutruzzolà, F.,
Di Lucente, C., Janson, G., Pascarella, S., Rinaldo, S.,
WORLD OF PLANT 14-3-3 PROTEINS S33
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Visconti, S., and Camoni, L. (2014) The phytotoxin
fusicoccin differently regulates 14-3-3 proteins as-
sociation to mode III targets, IUBMB Life, 66, 52-62,
https://doi.org/10.1002/iub.1239.
165. Saponaro, A., Porro, A., Chaves-Sanjuan, A.,
Nardini,M., Rauh,O., Thiel,G., and Moroni,A. (2017)
Fusicoccin activates KAT1 channels by stabilizing
their interaction with 14-3-3 proteins, Plant Cell, 29,
2570-2580, https://doi.org/10.1105/tpc.17.00375.
166. de Boer, A. H. (2024) The fusicoccin story revisited,
J. Exp. Bot., 75, 5531-5546, https://doi.org/10.1093/
jxb/erae300.
167. Sottocornola, B., Visconti, S., Orsi, S., Gazzarrini, S.,
Giacometti, S., Olivari, C., Camoni, L., Aducci, P.,
Marra, M., Abenavoli, A., Thiel, G., and Moroni, A.
(2006) The potassium channel KAT1 is activated by
plant and animal 14-3-3 proteins, J. Biol. Chem., 281,
35735-35741, https://doi.org/10.1074/jbc.M603361200.
168. Sottocornola,B., Gazzarrini,S., Olivari,C., Romani,G.,
Valbuzzi, P., Thiel, G., and Moroni, A. (2008) 14-3-3
proteins regulate the potassium channel KAT1 by
dual modes, Plant Biol., 10, 231-236, https://doi.org/
10.1111/j.1438-8677.2007.00028.x.
169. van Kleeff, P. J. M., Gao, J., Mol, S., Zwart, N.,
Zhang, H., Li, K. W., and de Boer, A. H. (2018) The
Arabidopsis GORK K
+
-channel is phosphorylated by
calcium-dependent protein kinase 21 (CPK21), which
in turn is activated by 14-3-3 proteins, Plant Physi-
ol. Biochem., 125, 219-231, https://doi.org/10.1016/
j.plaphy.2018.02.013.
170. Hedrich, R. (2012) Ion channels in plants, Physiol.
Rev., 92, 1777-1811, https://doi.org/10.1152/physrev.
00038.2011.
171. Jaspert, N., Throm, C., and Oecking, C. (2011) Arabi-
dopsis 14-3-3 proteins: fascinating and less fascinating
aspects, Front. Plant Sci., 2, 96, https://doi.org/10.3389/
fpls.2011.00096.
172. Prado, K., Cotelle, V., Li, G., Bellati, J., Tang, N.,
Tournaire- Roux, C., Martinière, A., Santoni, V., and
Maurela, C. (2019) Oscillating aquaporin phosphor-
ylation and 14-3-3 proteins mediate the circadian
regulation of leaf hydraulics, Plant Cell, 31, 417-429,
https://doi.org/10.1105/tpc.18.00804.
173. Prado,K., Boursiac,Y., Tournaire-Roux,C., Monneuse,
J.-M., Postaire, O., Da Ines, O., Schäffner, A. R.,
Hem, S., Santoni, V., and Maurel, C. (2013) Regula-
tion of Arabidopsis leaf hydraulics involves light-de-
pendent phosphorylation of aquaporins in veins,
Plant Cell, 25, 1029-1039, https://doi.org/10.1105/tpc.
112.108456.
174. Voelker, C., Gomez-Porras, J. L., Becker, D.,
Hamamoto, S., Uozumi, N., Gambale, F., Mueller-
Roeber, B., Czempinski, K., and Dreyer, I. (2010)
Roles of tandem-pore K
+
channels in plants – a puz-
zle still to be solved, Plant Biol., 12, 56-63, https://
doi.org/10.1111/j.1438-8677.2010.00353.x.
175. Latz,A., Becker,D., Hekman,M., Müller,T., Beyhl,D.,
Marten, I., Eing, C., Fischer, A., Dunkel, M., Bertl, A.,
Rapp, U.R., and Hedrich, R. (2007) TPK1, a Ca
2+
-reg-
ulated Arabidopsis vacuole two-pore K
+
channel is
activated by 14-3-3 proteins, Plant J., 52, 449-459,
https://doi.org/10.1111/j.1365-313X.2007.03255.x.
176. Sinnige, M. P., ten Hoopen, P., van den Wijngaard,
P. W. J., Roobeek, I., Schoonheim, P. J., Mol, J. N. M.,
and de Boer, A. H. (2005) The barley two-pore
K
+
-channel HvKCO1 interacts with 14-3-3 proteins in
an isoform specific manner, Plant Sci., 169, 612-619,
https://doi.org/10.1016/j.plantsci.2005.05.013.
177. Latz, A., Mehlmer, N., Zapf, S., Mueller, T. D.,
Wurzinger, B., Pfister, B., Csaszar, E., Hedrich, R.,
Teige, M., and Becker, D. (2013) Salt stress triggers
phosphorylation of the Arabidopsis vacuolar K
+
chan-
nel TPK1 by calcium-dependent protein kinases (CD-
PKs), Mol. Plant, 6, 1274-1289, https://doi.org/10.1093/
mp/sss158.
178. van den Wijngaard, P.W.J., Bunney, T.D., Roobeek,I.,
Schönknecht, G., and de Boer, A. H. (2001) Slow vac-
uolar channels from barley mesophyll cells are reg-
ulated by 14-3-3 proteins, FEBS Lett., 488, 100-104,
https://doi.org/10.1016/S0014-5793(00)02394-2.
179. Klychnikov, O., Li, K., Lill, H., and de Boer, A. (2007)
The V-ATPase from etiolated barley (Hordeum vul-
gare L.) shoots is activated by blue light and inter-
acts with 14-3-3 proteins, J. Exp. Bot., 58, 1013-1023,
https://doi.org/10.1093/jxb/erl261.
180. Cosse, M., and Seidel, T. (2021) Plant proton pumps
and cytosolic pH-homeostasis, Front. Plant Sci., 12,
672873, https://doi.org/10.3389/fpls.2021.672873.
181. Seidel,T. (2022) The plant V-ATPase, Front. Plant Sci.,
13, 931777, https://doi.org/10.3389/fpls.2022.931777.
182. Taoka, K., Kawahara, I., Shinya, S., Harada, K.,
Yamashita,E., Shimatani,Z., Furuita,K., Muranaka,T.,
Oyama, T., Terada, R., Nakagawa, A., Fujiwara, T.,
Tsuji,H., and Kojima,C. (2022) Multifunctional chem-
ical inhibitors of the florigen activation complex dis-
covered by structure-based high-throughput screen-
ing, Plant J., 112, 1337-1349, https://doi.org/10.1111/
tpj.16008.
183. Harada, K.ich., Furuita,K., Yamashita,E., Taoka, K.I.,
Tsuji, H., Fujiwara, T., Nakagawa, A., and Kojima, C.
(2022) Crystal structure of potato 14-3-3 protein St14f
revealed the importance of helix I in StFDL1 recogni-
tion, Sci. Rep., 12, 1-10, https://doi.org/10.1038/s41598-
022-15505-y.
184. Mishra, N.S., Tuteja,R., and Tuteja,N. (2006) Signal-
ing through MAP kinase networks in plants, Arch.
Biochem. Biophys., 452, 55-68, https://doi.org/10.1016/
j.abb.2006.05.001.
185. Oh, C.-S., Pedley, K.F., and Martin, G.B. (2010) Toma-
to 14-3-3 protein 7 positively regulates immunity-as-
sociated programmed Cell death by enhancing pro-
tein abundance and signaling ability of MAPKKK α,
SEDLOV, SLUCHANKOS34
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
Plant Cell, 22, 260-272, https://doi.org/10.1105/tpc.
109.070664.
186. Gao, Z., Zhang, D., Wang, X., Zhang, X., Wen, Z.,
Zhang, Q., Li, D., Dinesh-Kumar, S. P., and Zhang, Y.
(2022) Coat proteins of necroviruses target 14-3-3a to
subvert MAPKKKα-mediated antiviral immunity in
plants, Nat. Commun., 13, 716, https://doi.org/10.1038/
s41467-022-28395-5.
187. Teper, D., Salomon, D., Sunitha, S., Kim, J., Mudgett,
M. B., and Sessa, G. (2014) Xanthomonas euvesicato-
ria type III effector XopQ interacts with tomato and
pepper 14-3-3 isoforms to suppress effector-triggered
immunity, PlantJ., 77, 297-309, https://doi.org/10.1111/
tpj.12391.
188. Deb,S., Ghosh,P., Patel, H.K., and Sonti, R.V. (2020)
Interaction of the Xanthomonas effectors XopQ and
XopX results in induction of rice immune respons-
es, Plant J., 104, 332-350, https://doi.org/10.1111/
tpj.14924.
189. Taylor, K.W., Kim, J.-G., Su, X.B., Aakre, C.D., Roden,
J.A., Adams, C.M., and Mudgett, M.B. (2012) Tomato
TFT1 is required for PAMP-triggered immunity and
mutations that prevent T3S effector XopN from bind-
ing to TFT1 attenuate Xanthomonas virulence, PLoS
Pathog., 8, e1002768, https://doi.org/10.1371/journal.
ppat.1002768.
190. Lozano-Durán, R., Bourdais, G., He, S. Y., and
Robatzek, S. (2014) The bacterial effector HopM1
suppresses PAMP-triggered oxidative burst and sto-
matal immunity, New Phytol., 202, 259-269, https://
doi.org/10.1111/nph.12651.
191. Giska, F., Lichocka, M., Piechocki, M., Dadlez, M.,
Schmelzer,E., Hennig,J., and Krzymowska,M. (2013)
Phosphorylation of HopQ1, a type III effector from
Pseudomonas syringae, creates a binding site for Host
14-3-3 proteins, Plant Physiol., 161, 2049-2061, https://
doi.org/10.1104/pp.112.209023.
192. Evangelisti, E., Guyon, A., Shenhav, L., and
Schornack, S. (2023) FIRE mimics a 14-3-3-binding
motif to promote Phytophthora palmivora infection,
Mol. Plant Microbe Interact., 36, 315-322, https://
doi.org/10.1094/MPMI-12-22-0251-R.
193. Sheikh, A.H., Zacharia,I., Tabassum,N., Hirt,H., and
Ntoukakis, V. (2024) 14-3-3 proteins as a major hub
for plant immunity, Trends Plant Sci., 29, P1245-P1253,
https://doi.org/10.1016/j.tplants.2024.06.001.
194. Ji, H., Pardo, J. M., Batelli, G., Van Oosten, M. J.,
Bressan, R. A., and Li, X. (2013) The salt overly sen-
sitive (SOS) pathway: established and emerging
roles, Mol. Plant, 6, 275-286, https://doi.org/10.1093/
mp/sst017.
195. Qiu, Q. S., Guo, Y., Dietrich, M. A., Schumaker, K. S.,
and Zhu, J. K. (2002) Regulation of SOS1, a plasma
membrane Na
+
/H
+
exchanger in Arabidopsis thali-
ana, by SOS2 and SOS3, Proc. Natl. Acad. Sci. USA,
99, 8436-8441, https://doi.org/10.1073/pnas.122224699.
196. Christie, J.M. (2007) Phototropin blue-light receptors,
Annu. Rev. Plant Biol., 58, 21-45, https://doi.org/10.1146/
annurev.arplant.58.032806.103951.
197. Fankhauser, C., and Christie, J. M. (2015) Plant pho-
totropic growth, Curr. Biol., 25, R384-R389, https://
doi.org/10.1016/j.cub.2015.03.020.
198. Christie, J. M., Blackwood, L., Petersen, J., and
Sullivan, S. (2015) Plant flavoprotein photorecep-
tors, Plant Cell Physiol., 56, 401-413, https://doi.org/
10.1093/pcp/pcu196.
199. Sullivan,S., Thomson, C.E., Kaiserli,E., and Christie,
J. M. (2009) Interaction specificity of Arabidopsis
14-3-3 proteins with phototropin receptor kinases,
FEBS Lett., 583, 2187-2193, https://doi.org/10.1016/
j.febslet.2009.06.011.
200. Tseng, T.-S., Whippo, C., Hangarter, R. P., and Briggs,
W. R. (2012) The role of a 14-3-3 protein in stomatal
opening mediated by PHOT2 in Arabidopsis, Plant Cell,
24, 1114-1126, https://doi.org/10.1105/tpc.111.092130.
201. Christians, M.J., Gingerich, D.J., Hansen,M., Binder,
B.M., Kieber, J.J., and Vierstra, R.D. (2009) The BTB
ubiquitin ligases ETO1, EOL1 and EOL2 act collective-
ly to regulate ethylene biosynthesis in Arabidopsis by
controlling type-2 ACC synthase levels, PlantJ., 57, 332-
345, https://doi.org/10.1111/j.1365-313X.2008.03693.x.
202. Catalá, R., López-Cobollo, R., Mar Castellano, M.,
Angosto, T., Alonso, J. M., Ecker, J. R., and Salinas, J.
(2014) The Arabidopsis 14-3-3 protein RARE COLD
INDUCIBLE 1A links low-temperature response and
ethylene biosynthesis to regulate freezing tolerance
and cold acclimation, Plant Cell, 26, 3326-3342, https://
doi.org/10.1105/tpc.114.127605.
203. Mora-García, S., Vert, G., Yin, Y., Caño-Delgado, A.,
Cheong,H., and Chory,J. (2004) Nuclear protein phos-
phatases with Kelch-repeat domains modulate the re-
sponse to brassinosteroids in Arabidopsis, Genes Dev.,
18, 448-460, https://doi.org/10.1101/gad.1174204.
204. del Viso,F., Casaretto, J.A., and Quatrano, R.S. (2007)
14-3-3 Proteins are components of the transcription
complex of the ATEM1 promoter in Arabidopsis,
Planta, 227, 167-175, https://doi.org/10.1007/s00425-
007-0604-1.
205. Kawamoto,N., Sasabe,M., Endo,M., Machida,Y., and
Araki, T. (2015) Calcium-dependent protein kinases
responsible for the phosphorylation of a bZIP tran-
scription factor FD crucial for the florigen complex
formation, Sci. Rep., 5, 8341, https://doi.org/10.1038/
srep08341.
206. Guo, D., Yang, Z.-P., Li, H.-L., Wang, Y., Zhu, J.-H.,
and Peng, S.-Q. (2018) The 14-3-3 protein HbGF14a
interacts with a RING zinc finger protein to regu-
late expression of the rubber transferase gene in
Hevea brasiliensis, J. Exp. Bot., 69, 1903-1912, https://
doi.org/10.1093/jxb/ery049.
207. Yang, Z.-P., Li, H.-L., Guo, D., Tian, W.-M., and
Peng, S.-Q. (2012) Molecular characterization of a
WORLD OF PLANT 14-3-3 PROTEINS S35
BIOCHEMISTRY (Moscow) Vol. 90 Suppl. 1 2025
novel 14-3-3 protein gene (Hb14-3-3c) from Hevea
brasiliensis, Mol. Biol. Rep., 39, 4491-4497, https://
doi.org/10.1007/s11033-011-1239-7.
208. Yang, Z.-P., Li, H.-L., Guo,D., Tang,X., and Peng, S.-Q.
(2014) Identification and characterization of the 14-
3-3 gene family in Hevea brasiliensis, Plant Physi-
ol. Biochem., 80, 121-127, https://doi.org/10.1016/
j.plaphy.2014.03.034.
209. Sedlov, I. A., and Sluchanko, N. N. (2025) Biochemical
signatures strongly demarcate phylogenetic groups
of plant 14‐3‐3 isoforms, Plant J., 121, e70017, https://
doi.org/10.1111/tpj.70017.
210. Reiland, S., Messerli, G., Baerenfaller, K., Gerrits, B.,
Endler, A., Grossmann, J., Gruissem, W., and
Baginsky,S. (2009) Large-scale arabidopsis phosphopro-
teome profiling reveals novel chloroplast kinase sub-
strates and phosphorylation networks, Plant Physiol.,
150, 889-903, https://doi.org/10.1104/pp.109.138677.
211. Denison, F. C., Paul, A. L., Zupanska, A. K., and Ferl,
R.J. (2011) 14-3-3 proteins in plant physiology, Semin.
Cell Dev. Biol., 22, 720-727, https://doi.org/10.1016/
j.semcdb.2011.08.006.
212. Comparot,S., Lingiah,G., and Martin,T. (2003) Func-
tion and specificity of 14-3-3 proteins in the regula-
tion of carbohydrate and nitrogen metabolism, J.Exp.
Bot., 54, 595-604, https://doi.org/10.1093/jxb/erg057.
213. Yang, X., Wang, W., Coleman, M., Orgil, U., Feng, J.,
Ma, X., Ferl, R., Turner, J. G., and Xiao, S. (2009)
Arabidopsis 14-3-3 lambda is a positive regulator of
RPW8-mediated disease resistance, Plant J., 60, 539-
550, https://doi.org/10.1111/j.1365-313X.2009.03978.x.
214. Schoonheim, P. J., Veiga, H., da Costa Pereira, D.,
Friso, G., van Wijk, K. J., and de Boer, A. H. (2007)
A comprehensive analysis of the 14-3-3 Interactome
in barley leaves using a complementary proteomics
and two-hybrid approach, Plant Physiol., 143, 670-683,
https://doi.org/10.1104/pp.106.090159.
215. Kim, J., Lee, H., Lee, H. G., and Seo, P. J. (2021) Get
closer and make hotspots: liquid-liquid phase sep-
aration in plants, EMBO Rep., 22, e51656, https://
doi.org/10.15252/embr.202051656.
216. Wilson, R. S., Swatek, K. N., and Thelen, J. J. (2016)
Regulation of the Regulators: post-translational mod-
ifications, subcellular, and spatiotemporal distribu-
tion of plant 14-3-3 proteins, Front. Plant Sci., 7, 611,
https://doi.org/10.3389/fpls.2016.00611.
217. Bunney, T. D., Van Walraven, H. S., and De Boer,
A. H. (2001) 14-3-3 protein is a regulator of the mi-
tochondrial and chloroplast ATP synthase, Proc. Natl.
Acad. Sci. USA, 98, 4249-4254, https://doi.org/10.1073/
pnas.061437498.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in published
maps and institutional affiliations. AI tools may have
been used in the translation or editing of this article.