Content uploaded by Petronia Carillo
Author content
All content in this area was uploaded by Petronia Carillo on Mar 16, 2025
Content may be subject to copyright.
Content uploaded by Petronia Carillo
Author content
All content in this area was uploaded by Petronia Carillo on Mar 11, 2025
Content may be subject to copyright.
Review
Can biostimulants enhance plant resilience to heat and water stress in the
Mediterranean hotspot?
Petronia Carillo
Department of Environmental, Biological and Pharmaceutical Sciences and Technologies, University of Campania “Luigi Vanvitelli”, 81100 Caserta, Italy
ARTICLE INFO
Keywords:
Osmotic stress
Oxidative stress
Retrograde signalling
Osmolytes
heat shock factors (HSFs)
Heat shock proteins (HSPs)
Stress tolerance
ABSTRACT
Heat and water stress are imposing signicant constraints on agricultural systems, particularly in Mediterranean
regions experiencing prolonged droughts, rising temperatures, and increasing aridity. These abiotic stresses
trigger secondary effects, including osmotic and oxidative stress, simultaneously inuencing multiple plant traits.
Under drought conditions, stomatal closure limits CO₂ uptake, interfering with photosynthetic electron transport
and increasing the production of reactive oxygen species (ROS). Elevated ROS determine oxidative stress,
damaging cell membranes, causing genotoxicity, and disrupting key metabolic processes like nutrient transport,
cell division, and expansion. Plants activate natural defence mechanisms to counter these stresses, but these
responses are energetically costly. The diversion of carbon skeletons and energy from growth and biomass
accumulation to stress responses results in reduced yields, especially in key Mediterranean crops such as wheat,
tomato, grapevine, and olive trees, which are highly vulnerable to extreme climatic events. Biostimulants hold
signicant potential as an innovative approach to strengthening plants’ natural defences and enhancing their
capacity to endure heat and drought stress. By modulating stress-related pathways, enhancing antioxidant
defence mechanisms, and promoting the accumulation of osmolytes, these products help maintain water use
efciency (WUE), sustain photosynthetic activity, and reduce stress-induced yield losses. In areas where water
scarcity is a major limiting factor for agriculture, biostimulants offer a promising strategy to enhance plant
adaptation to increasingly unpredictable precipitation patterns and higher temperatures. Beyond their imme-
diate benets, biostimulants offer a sustainable solution for supporting crop productivity amidst climate change.
Further research into their biochemical, physiological, and metabolic impacts, specically focusing on Medi-
terranean cropping systems, will be essential to optimise their application and integrate them effectively into
modern, sustainable farming strategies.
1. Introduction
Plants are sessile organisms incapable of escaping from environ-
mental constraints. To cope with this limitation, they rely on
physiological responses that drive their developmental plasticity and
enable adaptation to varying abiotic factors, such as chemical and
physical conditions. These mechanisms support plant survival and are
not indicative of disease. However, when environmental changes are too
Abbreviations: Abscisic Acid, (ABA); Advanced Glycation End Products, (AGEs); Aminocyclopropane-1-Carboxylic Acid, (ACC); Aquaporins, (AQP); Arbuscular
Mycorrhizal, (AM); Ascophyllum nodosum Extract, (ANE); Ascorbate Peroxidase, (APX); Betaine Aldehyde Dehydrogenase, (BADH); Basic Helix-Loop-Helix, (bHLH);
Brassinosteroids, (BRs); Calcineurin B-Like Proteins, (CBLs); Calcium-Dependent Protein Kinases, (CDPKs); Carbon-Nitrogen Ratio, (C/N Ratio); Catalase, (CAT);
Chlorophyll, (Chl); Choline Monooxygenase, (CMO); Compatible Solutes, (CS); Cyclic Nucleotide-Dependent Calcium Channels, (CNGCs); Cyclin-Dependent Kinases,
(CDKs); Cyclin P2;1, (CYCP2;1); Cytokinins, (CK); Deg Proteases, (Deg); Dehydrins, (DHNs); Drought-Responsive Element Binding, (DREB); Electron Transport Chain,
(ETC); Ethylene, (ET); Field Capacity, (FC); Fucoidan, (Fu); Gamma-Aminobutyric Acid, (GABA); Gibberellins, (GA); Glutamine, (Gln); Glutathione, (GSH); Gluta-
thione Reductase, (GR); Glutathione S-Transferase, (GST); Heat Shock Factors, (HSFs); Heat Shock Proteins, (HSPs); Hydrogen Peroxide, (H₂O₂); Indole-3-Acetic Acid,
(IAA); Jasmonic Acid, (JA); Karrikins, (KAR); Late Embryogenesis Abundant, (LEA); Malondialdehyde, (MDA); Methylglyoxal, (MG); Methyl Jasmonate, (MeJA);
Monosilicic Acid, (MSA); Net photosynthesis, (Pn); Nitric Oxide, (NO); Phenylalanine Ammonia-Lyase, (PAL); Photosystem II, (PSII); Phytochrome Interacting
Factors, (PIF); Polyethylene Glycol, (PEG); Proline Dehydrogenase, (PDH); Reactive Carbonyl Species, (RCS); Reactive Oxygen Species, (ROS); Relative Water
Content, (RWC); Retrograde Signaling, (RS); Silicon, (Si); Stay-Green, (SGR); Stomatal Conductance, (gₛ); Strigolactones, (SLs); Superoxide, (O₂•⁻); Superoxide
Dismutase, (SOD); Transcription Factors, (TFs); Tricarboxylic Acid Cycle, (TCA); Ulva lactuca Extract, (ULE); Water Use Efciency, (WUE); Zinc, (Zn).
E-mail address: petronia.carillo@unicampania.it.
Contents lists available at ScienceDirect
Plant Stress
journal homepage: www.sciencedirect.com/journal/plant-stress
https://doi.org/10.1016/j.stress.2025.100802
Received 9 January 2025; Received in revised form 19 February 2025; Accepted 10 March 2025
Plant Stress 16 (2025) 100802
Available online 11 March 2025
2667-064X/© 2025 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (
http://creativecommons.org/licenses/by-
nc-nd/4.0/ ).
rapid or intense, plants perceive them as stressors. In recent years,
exacerbating environmental conditions and meteorological phenomena
have led to heightened intensity and frequency of both single and
multiple abiotic stress phenomena for plants, signicantly reducing
global agricultural production (Benitez-Alfonso et al., 2023). High
temperatures and drought are the primary environmental stresses that
affect plant growth and productivity by directly impairing physiological
functions, altering metabolic pathways, and reducing resource avail-
ability. These stressors also damage natural resources, crops, livestock,
and livelihoods. Drought stress limits water availability, restricting cell
expansion, reducing stomatal conductance, and compromising photo-
synthetic efciency, thus leading to lower biomass accumulation. Pro-
longed water decits cause oxidative stress, induce hormonal
imbalances, and alter root architecture, affecting plant development at
all growth stages (Osakabe et al., 2014). Conversely, heat stress accel-
erates metabolic processes, increases transpiration rates, and disrupts
cellular homeostasis. Extreme temperatures denature proteins, impair
membrane stability, and inhibit enzymatic functions essential for carbon
xation and energy production (dos Santos et al., 2022). When drought
and heat stress co-occur, their combined impact is often more severe,
exacerbating physiological and biochemical disruptions beyond the ef-
fects of each stress alone (Sato et al., 2024). In the Horn of Africa, an
extreme drought, the most severe in forty years, has persisted, with
nearly four years of no rainfall (UNOCHA, 2023; WEF, 2022). This crisis
has resulted in widespread food shortages, malnutrition, and displace-
ment, severely impacting livelihoods and stability across the affected
areas. Here, the control of water could lead to conicts as nations and
groups compete for stable water supplies amid shortages, pollution, and
climate change (Syed, 2020). Similarly, the Mediterranean basin faces a
progressive decline in precipitation and increasing temperatures,
intensifying water scarcity and placing agricultural systems under se-
vere pressure. This area has been identied as one of the most vulnerable
regions to climate change, with annual temperatures rising approxi-
mately 20% faster than the global average, with rainfall patterns
becoming increasingly erratic (Lazoglou et al., 2024; Le Monde, 2023).
Between January and September 2023, global temperatures rose by 1.45
◦C above pre-industrial levels, making it the hottest year recorded at the
time (WMO, 2023). Summer 2023 was the warmest globally, with
Europe experiencing its fth hottest summer (Copernicus, 2025). This
warming trend intensied in 2024, with temperatures rising 1.55 ◦C
above pre-industrial levels, marking another record-breaking summer
(WMO, 2025). These climatic shifts are already severely impacting
Mediterranean agriculture, where rising temperatures and prolonged
droughts have led to signicant declines in key crop yields. By the end of
2023, the situation in Spain appeared so critical that both the New
Zealand Ministry of Foreign Affairs and Trade (MFAT, 2023) and the
Dutch Ministry of Agriculture, Nature and Food Quality (Agroberichten
Buitenland, 2024) published reports highlighting the severe conse-
quences of climate change on Spanish agriculture. These analyses
highlighted the persistent drought, record-high temperatures, and water
shortages that caused an unprecedented decline in key crops
(Copernicus, 2025). Particularly Spain cereals production fell to 11.9 Mt
in 2023, 38% and decreased from 19.3 Mt in 2022 and 25.5 Mt in 2021,
making it one of the most severe drought-driven declines in recent de-
cades. The impact was particularly acute in Castilla y Le´
on, the country’s
primary wheat-producing region, where yields dropped 40% below the
ve-year average. Common wheat and spelt production declined to 3.6
Mt in 2023, compared to 5.8 Mt in 2022 and 7.8 Mt in 2021, under-
scoring the severe water shortages and extreme heat stress affecting key
growth stages (EC, 2025). The sharp decline in Spain’s grain production
and disruptions in Ukrainian grain exports worsened global food
shortages and added to market instability. Despite these alarming g-
ures, early 2024 data suggest a partial recovery for most sectors, except
for durum wheat and common wheat, which remain signicantly below
pre-drought levels. However, concerns persist regarding the long-term
impacts of climate change, as water scarcity continues to challenge
Spain’s agricultural resilience. Indeed the impact of climate change on
Mediterranean agriculture is further exacerbated by socio-economic
factors, including a population that has more than tripled in the last
70 years, causing a rise of atmospheric CO
2
levels from below 300 ppm
to 426.9 ppm as of May 2024 (NOAA, 2024). Recent reports indicate
that without signicant mitigation efforts, global temperatures are
projected to increase by about 3.1 ◦C by the end of the century,
threatening ecosystems, human health, and economies on a global scale
(Carbon Brief, 2024). In addition to temperature increases, reduced
water availability is becoming a major constraint on Mediterranean crop
production. Projections indicate a 4% to 22% reduction in annual pre-
cipitation, with southern Europe potentially seeing up to a 30% decrease
in winter precipitation (Kostopoulou and Giannakopoulos, 2023; Phi-
landras et al., 2011; IPCC 2022). In European Mediterranean countries,
the intensication and prolonged duration of this water scarcity have
elevated water to a critical resource. Water scarcity is primarily driven
by demand and consumption, with agriculture accounting for about
70%, industry 19%, and municipal use 11%. Italy is among the European
countries experiencing medium to extremely high water stress, with the
Emilia-Romagna region acting as a climatic demarcation line on the
peninsula. The ood in May 2023 was a pivotal event that marked the
end of the drought in Northern Italy and shifted the focus of water
scarcity towards the southern regions and major islands (Copernicus,
2023; Euronews, 2023). However, by late 2024, water scarcity had
intensied in southern regions and major islands. Data from Greenpeace
and the National Research Council (CNR) indicate that approximately
69% of Sicily’s territory and 47% of Calabria’s were affected by extreme
drought conditions (Icona Clima, 2023). These conditions have severely
impacted water resources and local communities, underscoring the ur-
gent need for comprehensive water management strategies in these
areas (Blue Community, 2024; The Independent, 2023). Crop yields are
expected to decrease by roughly 3–5% with every 1 ◦C rise in temper-
ature. This reduction is primarily due to the shortening of the crop
growing season and increased heat and water stress during critical
growth stages. These impacts are expected to be more severe in southern
Mediterranean areas, where higher temperatures and reduced rainfall
will likely exacerbate water scarcity and crop stress (Grasso and Feola,
2012; IPCC, 2022). Consequently, the Mediterranean agricultural sector
is under pressure, requiring innovative strategies to reduce the harmful
effects of climate change on food security and crop yields. One prom-
ising approach is using biostimulants, which have been shown to
enhance plant resilience to abiotic stresses by modulating key physio-
logical and biochemical pathways. Derived from natural substances or
microorganisms, these products stimulate plant physiological processes,
improving nutrient utilisation, stress tolerance under adverse conditions
such as drought and salinity, and overall crop productivity and quality
(Carillo et al., 2025). The European Regulation (EU) 2019/1009 denes
biostimulants as products that “stimulate plant nutrition processes
independently of the product’s nutrient content with the sole aim of
improving one or more of the following characteristics of the plant or the
plant rhizosphere: (a) nutrient use efciency; (b) tolerance to abiotic
stress; (c) quality traits; or (d) availability of conned nutrients in the
soil or rhizosphere.” (EU, 2019). Biostimulants improve water use ef-
ciency, protect photosynthetic processes, and mitigate oxidative stress,
thereby supporting plant growth under adverse conditions (Carillo,
Kyratzis, et al., 2020; Fernandes et al., 2022; Gedeon et al., 2022; Iacuzzi
et al., 2024; Rouphael et al., 2023; Spada et al., 2024). However, despite
increasing research on their benets, further studies are required to
optimise their application and understand their long-term effects in
Mediterranean cropping systems (Fusco et al., 2022; Jim´
enez-Arias
et al., 2022). This review aims to explore the impact of heat and drought
stress on plant physiology, with a specic focus on Mediterranean
agriculture, and to assess the potential of biostimulants as a sustainable
tool to mitigate climate-induced yield losses.
P. Carillo
Plant Stress 16 (2025) 100802
2
2. Effects of water and heat stress on plant growth and
development
The challenges for plants under drought conditions begin as early as
the seed stage. Seed hydration, a critical step for activating key enzymes
essential for germination, is signicantly hindered under water stress.
Drought-induced water shortages reduce these enzymes’ activity,
resulting in poor germination, difculty in seedling establishment, and a
loss of seedling strength (Biju et al., 2017). Gibberellic acid (GA), a
primary hormone promoting seed germination, plays a crucial role in
activating hydrolytic enzymes that break down seed reserves to support
growth. However, its function is antagonised by abscisic acid (ABA),
which inhibits germination by suppressing GA activity (Saha et al.,
2022; Tong et al., 2017). Accordingly, water stress induced by drought
or polyethylene glycol (PEG-6000), can signicantly delay or inhibit
seed germination by reducing water potential, thereby limiting the
essential water uptake required for seed hydration and emergence. This
stress elevates ROS levels, triggering calcium ion (Ca²⁺) cascades and
ABA accumulation, prolonging seed dormancy and hindering seedling
establishment(Oracz and Karpi´
nski, 2016). For instance, the expression
of OsNCED3, a key gene in the biosynthesis of ABA in rice, is signi-
cantly upregulated under drought stress, resulting in increased ABA
accumulation (Liu et al., 2019). Additionally, drought stress lowers
indole-3-acetic acid (IAA) and cytokinins (CK) levels, further hindering
seedling establishment (Prerostova et al., 2018). Under moderate water
stress, seedlings may not exhibit signicant changes in overall size;
however, an increase in root growth often occurs as a compensatory
response to optimise water uptake. This adaptive mechanism enhances
the plant’s ability to access deeper soil moisture, thereby mitigating the
effects of limited water availability (Thomas et al., 2024). Conversely,
severe or prolonged water stress adversely affects shoot and root growth.
This intense stress impairs photosynthesis, reducing the production and
transport of photosynthates, and consequently diminishes root biomass.
The decrease in root mass further limits water uptake, compromising the
development of shoots and roots (Osakabe et al., 2014). Accordingly, a
study on basil (Ocimum basilicum cv. Genovese) demonstrated that under
severe drought conditions (30% soil water capacity), plants exhibited a
50% reduction in both fresh and dry yield, along with shorter stature
and narrower canopies (Mulugeta and Rad´
acsi, 2022). Similarly,
research on tomato (Solanum lycopersicum cv. Marmande) under drought
stress (50% eld capacity, FC) revealed signicant reductions in growth
and physiological parameters, including a 43% decrease in shoot weight
and 17% in total chlorophyll content, along with declines in key hor-
mones like GA (93%) and salicylic acid (SA) (82%) (Turan et al., 2023).
Similar effects have been observed in other Mediterranean staple crops.
In durum wheat (Triticum durum), drought stress reduces grain yield and
quality by limiting carbon assimilation and disrupting grain lling
(Kakanur Jagadeesha et al., 2024; KiliÇ and Tacettin, 2010; Moral et al.,
2003). In grapevine (Vitis vinifera), water decit affects berry develop-
ment, modulating secondary metabolism and altering polyphenol con-
tent, with signicant implications for wine production (Cataldo et al.,
2023; Chac´
on-Vozmediano et al., 2021; Lovisolo et al., 2010; Savoi
et al., 2020). Additionally, olive trees (Olea europaea) exhibit high
drought tolerance, mainly through stomatal regulation and osmopro-
tectant accumulation (Sofo et al., 2004), yet prolonged stress can impair
fruit yield and oil quality (Brito et al., 2019; Gholami et al., 2022; Parri
et al., 2023). These ndings emphasise the necessity of implementing
water-saving strategies, such as decit irrigation and mulching, to sus-
tain productivity in drought-prone regions while optimising water use
efciency (Bekele and Asefa, 2023). High temperatures further impede
seed germination through cellular damage, reduced water availability,
increased respiration rates, protein denaturation, and desiccation, ulti-
mately leading to germination failure. For instance, seed germination
rates can decrease by approximately 40% at 32 ◦C compared to optimal
conditions at 24 ◦C (Ghaleb et al., 2022). Seed problems can also stem
from the parent plant’s exposure to heat stress. Moderately high
temperatures impair grain lling and reduce grain size, signicantly
decreasing rice yield and quality. Prolonged exposure to severe heat
stress (39 ◦C) amplies these impacts, causing grain malformation and
further compromising crop productivity (Begcy et al., 2018).
At moderately high temperatures (27–32 ◦C), newly germinated
seedlings undergo thermomorphogenesis, displaying elongation of the
hypocotyl or epicotyl and upward leaf bending (hyponasty or thermot-
ropism). These adaptations help seedlings distance themselves from the
heat-retaining ground and enhance cooling by increasing exposure to air
currents. Such moderate heat stress accelerates developmental transi-
tions and structural changes without causing plant death (Vu, Gevaert,
et al., 2019). The stem elongation and leaf hyponasty associated with
heat stress depend on the basic helix-loop-helix (bHLH) transcription
regulators Phytochrome Interacting Factor 4 (PIF4) and PIF7 (Koini
et al., 2009). These regulators act through IAA, which is transported
from the leaves to the petiole and then reaches the hypocotyl or epicotyl
under high-temperature conditions. In conjunction with brassinoste-
roids (BRs), IAA promotes growth by elongation and/or expansion. This
effect is enhanced by stabilising hormone receptors and increased hor-
mone transport at elevated temperatures (Nemhauser et al., 2004; Vu,
Xu, et al., 2019). In addition, the temperature-sensitive PIN-LIKES (PILS)
proteins in the endoplasmic reticulum (ER) play a role in intracellular
auxin distribution. PILS6, in particular, decreases under heat shock
conditions, enabling subcellular auxin redistribution and enhancing its
signalling capacity, which promotes thermomorphogenic growth (Li
et al., 2021). In contrast, high temperatures signicantly impact the
germination of tomato seeds and the development of plants. Toki´
c et al.
(2023) investigated the effect of elevated temperatures on tomato
seedlings under conditions mimicking frequent summer heat in Medi-
terranean semi-arid and arid regions. Seedlings were subjected to tem-
peratures of 37 ◦C and extreme heat conditions of 45 ◦C. Primary root
length was affected by both heat stresses, whereas lateral root devel-
opment was notably suppressed only at 37 ◦C. The 37 ◦C treatment also
led to a notable increase in the ethylene precursor 1-aminocyclopropa-
ne-1-carboxylic acid (ACC), likely in modifying root architecture. In
contrast, the treatment at 45 ◦C triggered pronounced physiological and
morphological damage, including chlorosis, leaf wilting, and stem
bending, in both young and mature plants. These severe phenotypic
changes were accompanied by accumulated stress markers such as
proline, malondialdehyde (MDA), and heat shock protein (HSP) 90.
Additionally, heat stress markedly impacted the expression of tran-
scription factors (TFs) related to stress responses, with DREB1 emerging
as a consistent heat stress marker, highlighting its potential for devel-
oping heat-tolerant crops (Toki´
c et al., 2023). In chickpeas (Cicer arie-
tinum), high temperatures during owering impair pod set and seed
development, leading to signicant yield losses (Jumrani and Bhatia,
2014). Lentil (Lens culinaris) shows a similar sensitivity, with heat stress
reducing seed viability and germination rates (Sita et al., 2017). In
durum wheat, elevated temperatures accelerate grain lling but
compromise starch accumulation, affecting pasta-making quality
(Labuschagne et al., 2009; Sabella et al., 2020). Therefore, extreme heat
stress (temperatures above 40 ◦C) can trigger rapid damage to cells and
result in plant death. However, prior exposure to sub-lethal high tem-
peratures (below 37 ◦C) can trigger acquired thermotolerance,
enhancing a plant’s ability to survive subsequent extreme heat events
(Vu, Xu, et al., 2019). This emphasises immediate and acclimatory heat
response mechanisms in plant resilience, including thermomorpho-
genesis and heat stress responses (HSR). Recent studies reveal that
plants signicantly increase transpiration rates under high-temperature
stress, even in dry conditions. This response is driven by physical
changes, such as reduced water viscosity and the melting of cuticular
waxes, alongside physiological adjustments, including enhanced aqua-
porin activity, potassium-mediated transmembrane water transport, and
altered membrane uidity. These mechanisms enable evaporative
cooling, reducing leaf temperatures by up to 14 ◦C, helping maintain
optimal photosynthesis conditions. However, the increased water loss
P. Carillo
Plant Stress 16 (2025) 100802
3
can lead to decits that adversely impact critical stages of seed devel-
opment and lling, revealing a compromise between cooling and water
conservation under heat stress (Sadok et al., 2021).
3. Understanding plant stress resistance mechanisms as a key to
adaptation
High temperature and drought stress induce two major secondary
stresses: osmotic and oxidative stress. These stresses cause pleiotropic
effects, simultaneously impacting multiple plant traits. Stomatal closure
reduces CO₂ uptake, disrupting photosynthetic electron transport and
elevating ROS levels. This oxidative stress causes genotoxicity, damages
cell membranes, and disrupts metabolic processes, hindering nutrient
transport, cell division, and expansion. These disruptions culminate in
reduced plant growth and yield, emphasising the complex interplay of
stress factors that affect crop productivity (Cruz de Carvalho, 2008;
Pandey et al., 2015).
Increased membrane uidity at high temperatures and impaired
electron transport under drought generate retrograde signals that
modulate nuclear gene expression via calcium uxes. Dysfunctions in
these transport chains lead to ROS accumulation, mainly hydrogen
peroxide (H₂O₂), which acts as both a stress marker and a regulatory
signal for adaptive responses (Foyer and Hanke, 2022), While excessive
ROS cause cellular damage, moderate concentrations integrate redox
uctuations with environmental cues, linking oxidative stress with
stress-induced senescence (Carillo and Ferrante, 2025; Savchenko and
Tikhonov, 2021). Similarly, ROS generated through ABA-mediated sig-
nalling pathways play a vital role in stress adaptationby inducing H₂O₂
production via NADPH oxidase, positioning it as a key effector in
regulating stress responses. This signaling cascade activates genes
involved in osmolytes synthesis, such as proline and glycine betaine
(GB), vital for osmotic adjustment, water retention, and oxidative
damage mitigation under drought and/or heat stress. Osmotic and ionic
stresses also trigger the generation of reactive carbonyl species (RCS),
such as methylglyoxal (MG). MG plays a dual role: at high concentra-
tions, it is cytotoxic and leads to the formation of Advanced Glycation
End Products (AGEs), which impair protein and DNA functions, while at
low concentrations, it acts as a signalling molecule ne-tuning retro-
grade responses (Sun et al., 2024). MG detoxication is managed by the
glyoxalase (GLO) system, which primarily involves
glutathione-dependent glyoxalase I and II, converting MG into less toxic
products like d-lactate, thereby preventing AGE accumulation and pre-
serving cellular homeostasis(Hoque et al., 2016). In an experiment by
Wang et al. (2020), Triticum aestivum plants subjected to simulated
drought (soil water potential of −0.8 to −1 MPa for 5 days) and heat
stress (daytime temperatures >35 ◦C) during the grain-lling stage
exhibited impaired stress responses when ABA biosynthesis was
inhibited using uridone (1 µM). This inhibition reduced ABA and H₂O₂
levels, disrupting mechanisms essential for stress resilience, ultimately
causing signicant declines in leaf water potential and grain yield.
Conversely, scavenging H₂O₂ with dimethylthiourea (5 mM) modied
downstream responses without affecting ABA synthesis, emphasizing
the indispensable role of H₂O₂ as a downstream mediator in ABA-driven
adaptation to stress. ROS also activate calcium signalling and interact
with other secondary messengers, such as nitric oxide (NO), enhancing
the plant’s capacity to sense and react to environmental stress (Sandalio
et al., 2023). Calcium concentrations vary signicantly across cellular
compartments: approximately 10⁻³ M outside the cell, around 100
micromolar in the endoplasmic reticulum, and 0.5–200 nM in the
cytosol under resting conditions. However, cytosolic calcium levels rise
approximately tenfold upon stimulation, reaching 1–2 micromolar. High
temperatures further increase plasma membrane uidity, activating
heat sensors that facilitate transient Ca²⁺ inux from the apoplast into
the cytosol, initiating signalling events. Takahashi et al. (2020) noted
that during drought stress conditions, ROS and Ca²⁺ waves serve as
mobile signals enabling communication between tissues and over long
distances, vital for whole-plant drought stress resistance. In particular,
root-to-shoot signallings, such as hydraulic pressure and ROS/Ca²⁺
waves, in addition to peptide signals, are critical under dehydration
stress. Shoot tissues mediate stomatal closure, ensuring water conser-
vation during stress. Locally, these signals, peptides, and ABA regulate
stomatal function to optimise plant WUE (Takahashi et al., 2020).
Beyond local responses, they also trigger systemic thermotolerance and
drought resistance pathways by stabilizing membrane-enhancing anti-
oxidant defences, promoting the accumulation of osmolytes, and upre-
gulating heat shock proteins (HSPs). This interplay ne-tunes stress
responses, ensuring rapid response to heat and drought stress
Exposure to high temperatures (up to 37 ◦C) and/or mild water stress
can trigger plant stress tolerance responses. One response is the heat
shock response (HSR), a signalling cascade that activates heat shock
factors (HSFs). These TFs regulate the gene expression prole associated
with heat stress, initiating protective responses such as synthesising
HSPs. The primary temperature sensor in plants appears to be an in-
crease in plasma membrane uidity, which activates cyclic nucleotide-
dependent calcium channels. This Ca²⁺ inux during heat shock trig-
gers intracellular signalling cascades, synergistic and/or interrelated
with redox changes in chloroplasts, mediated by the H₂O₂ molecules,
that activate HSFA1 (Saidi et al., 2011). This latter is a key regulator that
drives the expression of HSP70 and HSP90, addressing protein mis-
folding and aggregation caused by heat stress (Dickinson et al., 2018).
HSPs have highly conserved sequences and act as molecular chaperones
to stabilise proteins and avoid misfolding or aggregation, particularly
under thermal and drought stress conditions. Correct protein folding
(secondary and tertiary structures) and the assembly of multisubunit
protein complexes (quaternary structures) are essential for protein
function (Maytin, 1995; Wang et al., 2004). In Arabidopsis thaliana,
diurnal variations in thermotolerance align with increased HSP70
expression during the day, inuenced by chloroplast-derived redox
signals. HSFA1 further promotes HSF2 transcription, forming an
HSFA1-HSFA2 superactivator complex that amplies HSPs production
and establishes acquired thermotolerance. This intricate network of HSF
activation, redox signalling, and HSP synthesis allows plants to with-
stand high temperatures and adapt to diverse environmental stresses
effectively. Recent research highlights the functions of hormones like
BRs and ABA in enhancing heat stress tolerance by modulating the ac-
tivity of HSFs. Furthermore, recent studies reveal that various cellular
compartments detect heat stress, including the nucleus, where mecha-
nisms such as nuclear condensate formation through liquid-liquid phase
separation are crucial for coordinating stress responses (Dündar et al.,
2024).
High temperatures and drought often result in excessive evapo-
transpiration, inducing osmotic stress in plants. To address this, the
accumulation of osmolytes is essential for fostering tolerance to these
abiotic stresses. In cereals, osmolytes such as proline and GB function as
osmoregulators, maintaining cellular water balance under stress. These
molecules, along with antioxidants like ascorbic acid (AA), glutathione
(GSH), and various antioxidant enzymes, stabilise membranes and
proteins, maintain redox balance, and strengthen the plant’s capacity to
withstand stress. Protecting cellular structures and functions makes
these compounds essential for plants to thrive under adverse conditions.
However, osmoregulation and antioxidant defences come at a high en-
ergetic cost, requiring plants to use approximately 50–70 moles of ATP
to synthesise each mole of osmolytes or protective metabolites (Raven,
1985). This high energy demand diverts carbon skeletons and energy
from growth-related pathways, reducing biomass accumulation and
lower yields. This trade-off underscores the substantial impact of stress
responses on plant productivity, as resources are reallocated to prioritise
survival over growth during abiotic stress conditions.
4. Integrating agronomic approaches for climate adaptation
Given the intrinsic vulnerability of agricultural productivity to
P. Carillo
Plant Stress 16 (2025) 100802
4
climate change, implementing adaptive agronomic strategies is essential
to ensure long-term yield stability and resource-use efciency. Among
the most effective approaches, sowing date modications, decit irri-
gation techniques, shading systems, and soil conservation practices have
been identied as key measures to mitigate the adverse effects of rising
temperatures and irregular water availability (Agho et al., 2024).
However, these strategies must be carefully tailored to specic envi-
ronmental conditions, crop physiology, and local agronomic constraints
to maximise their effectiveness and minimise trade-offs (Minoli et al.,
2022).
Optimising sowing dates is one of the primary adaptation strategies
available to farmers in semi-arid Mediterranean environments, where
temperature and rainfall patterns dictate the success of crop establish-
ment, development, and yield formation. Bread wheat (Triticum aestivum
L.), phenological responses to thermal regimes are particularly signi-
cant, as elevated temperatures accelerate growth cycles, shortening the
grain-lling period and compromising both yield and quality (Hossain
and Teixeira Da Silva, 2012). Adjusting sowing schedules can help
synchronise critical developmental phases with favourable climatic
conditions, mitigating heat stress during anthesis and ensuring adequate
water availability during grain lling (Agho et al., 2024). However, the
success of this approach is closely tied to cultivar selection, as different
wheat varieties exhibit distinct vernalisation requirements and thermal
thresholds that determine their adaptation potential under shifting
climate conditions (Gouache et al., 2012). Recent advances in precision
agriculture have further rened the ability to optimise sowing dates,
with remote sensing tools such as the Normalized Difference Vegetation
Index (NDVI) and the Normalized Difference Red Edge (NDRE) enabling
real-time assessments of plant water status and nutrient uptake
(Allende-Montalban et al., 2024). By integrating these technologies with
predictive climate models, farmers can dynamically adjust management
practices, minimising exposure to climatic risks while enhancing
resource-use efciency. Nevertheless, modifying sowing dates alone
does not fully counteract the challenges posed by increasing tempera-
tures and erratic precipitation. Extreme weather events, unexpected
heat waves, and prolonged dry spells continue to threaten crop pro-
ductivity, requiring additional agronomic measures to bolster resilience
(Georgopoulou et al., 2024).
Decit irrigation (DI) is another widely adopted agronomic strategy
for optimising WUE without excessive yield penalties in Mediterranean
areas where water scarcity represents a fundamental constraint to yield
stability (Fereres and Soriano, 2006). Among the most prominent DI
techniques, sustained decit irrigation (SDI) involves a uniform reduc-
tion in water supply throughout the growing season, whereas regulated
decit irrigation (RDI) strategically withholds water during non-critical
phenological stages to maintain productivity while conserving resources
(Burato et al., 2024; Galindo et al., 2018; Zarrouk et al., 2016). Partial
root-zone drying (PRD) alternates irrigation between root zones, stim-
ulating ABA-mediated or independent stomatal regulation to enhance
drought tolerance without entirely restricting water uptake (Jovanovic
and Stikic, 2018; Kang and Zhang, 2004; P´
erez-P´
erez et al., 2020). These
methods have demonstrated considerable potential in Mediterranean
crops, such as tomato, olive, lemon, grapevine, and pomegranate,
improving fruit composition while reducing overall water consumption
(Gil et al., 2018; Giuliani et al., 2017; P´
erez-P´
erez et al., 2012; Wahbi
et al., 2005). However, the practical implementation of DI strategies is
often hindered by site-specic constraints, including soil heterogeneity,
unpredictable rainfall patterns, and the difculty of precisely moni-
toring soil moisture levels (O’Donnell et al., 2023). Furthermore, con-
cerns regarding long-term soil salinisation and nutrient leaching remain
insufcient, emphasising the need for a more integrated approach
combining water management with complementary agronomic prac-
tices (Aragü´
es et al., 2015).
Soil conservation strategies are also crucial in enhancing crop resil-
ience to climate change, particularly in rainfed Mediterranean systems
where water retention is a limiting factor. Among these, straw mulching
and reduced tillage have proven highly effective in mitigating soil
evaporation, stabilising temperature uctuations, and improving soil
structure. By forming a protective layer over the soil surface, mulching
minimises direct exposure to solar radiation, reducing moisture loss
while fostering microbial activity and nutrient cycling (Chen et al.,
2007). In parallel, reduced tillage maintains soil integrity, enhancing
aggregate stability and organic matter content while limiting mechani-
cal disturbance to microbial communities (Zhang et al., 2016). Inte-
grating no-tillage (NT) with low-emission fertilisation further improved
durum wheat yield (+15%) and grain quality in dry years, enhancing
nutrient use efciency. This combination also maintained crop perfor-
mance and increased economic returns, highlighting its potential for
sustainable wheat production under water-limited conditions (De Santis
et al., 2024). However, the impact of NT on crop yield in semi-arid re-
gions remains variable. A study in southern Italy found that NT
improved soil water retention and boosted yields in drier years, partic-
ularly in Foggia. At the same time, in wetter conditions, conventional
tillage (CT) led to higher grain protein content, likely due to different
water dynamics (De Vita et al., 2007). This means their effectiveness is
highly dependent on site-specic conditions, with some studies report-
ing initial declines in productivity due to transitional shifts in soil
structure and microbial dynamics. Beyond soil and irrigation manage-
ment, adopting shading systems has gained increasing attention as an
effective means of mitigating heat and radiation stress, particularly in
high-temperature environments where excessive solar exposure can
compromise physiological performance. Shading nets, which vary in
mechanical and optical properties, modulate microclimatic conditions
by reducing radiation intensity, stabilising temperature uctuations,
and altering the spectral composition of transmitted light (Boini et al.,
2021; Caruso et al., 2020; Statuto et al., 2019). Traditional shading
materials, such as black or grey nets, primarily function by lowering
light intensity, while photo-selective nets have been designed to
enhance diffuse radiation, optimising photosynthesis while minimising
photoinhibition (Formisano, 2022). Shading nets have been widely
studied in heat-sensitive horticultural crops, including bell pepper, let-
tuce, and rocket, where they have been shown to increase leaf expan-
sion, regulate stomatal activity, and improve nutrient uptake (Caruso
et al., 2020; Díaz-P´
erez and St. John, 2019; Formisano et al., 2021).
Additionally, shading contributes to post-harvest quality by maintaining
higher levels of bioactive compounds, such as avonoids and glucosi-
nolates, in leafy vegetables (Caruso et al., 2020; Jim´
enez-Viveros and
Valiente-Banuet, 2023). However, shading can also reduce carbohydrate
accumulation due to lower photosynthetic rates under diminished light
availability. Therefore, the balance between heat stress mitigation and
optimal light interception must be carefully managed to ensure that
shading does not inadvertently suppress yield potential (Gallo et al.,
2024).
While these strategies provide substantial benets, their effective-
ness depends on environmental conditions, crop physiology, and farm-
level management, requiring integration into a broader framework of
sustainable intensication. The combined implementation of these
adaptive practices enhances resilience to climate-induced stressors.
However, no single approach can fully mitigate the multifaceted chal-
lenges of climate change, suggesting that further integration and
renement may be necessary to maximize their impact. Emerging
research recognizes biostimulants as a promising complement to agro-
nomic strategies, with potential for enhancing crop resilience and sus-
tainability. Their effectiveness may further improve through integration
with existing practices and adaptive approaches (Franzoni et al., 2022;
Rouphael and Colla, 2020).
5. Biostimulants as a strategy to alleviate heat and drought
stress in plants
Biostimulants are increasingly acknowledged as a crucial asset in
modern agriculture, promoting crop growth, resilience, and quality
P. Carillo
Plant Stress 16 (2025) 100802
5
while supporting sustainable farming practices. Derived from natural
substances or microorganisms, these products enhance plant natural
processes, improving nutrient use efciency (NUE), stress tolerance (e.
g., drought and salinity), and overall plant health and productivity. The
European Biostimulants Industry Council (EBIC, https://biostimulants.
eu/), representing the plant biostimulants industry in policy and prac-
tice, has been instrumental in advocating for their commercial potential
and economic value. At the same time, research into biostimulants over
the past 15 years has played a crucial role in demonstrating their ef-
cacy, mode and mechanisms of action. Numerous European research
programs have provided signicant funding and resources to advance
the understanding and application of biostimulants. In collaboration
with industry stakeholders, these initiatives have driven innovation and
expanded the adoption of biostimulants across diverse agricultural
systems. On the policy front, the European Union’s Regulation 2019/
1009 has established a global benchmark by dening biostimulants as
products that enhance plant growth by supporting natural processes
without directly supplying nutrients. This regulation has claried their
classication and use while ensuring compliance with strict safety and
environmental standards. Furthermore, the European Innovation Part-
nership for Agricultural Productivity and Sustainability (EIP-AGRI),
launched by the European Commission in 2012, emphasizes the stra-
tegic importance of biostimulants in addressing sustainability challenges
and fostering innovation within agriculture (EC, 2023). Together, these
efforts reect a unied momentum from research, industry, and policy
to position biostimulants as a cornerstone of sustainable farming. The
following sections will examine specic biostimulant products that have
demonstrated efcacy against heat and drought stress, showcasing their
potential to enhance plant resilience under challenging environmental
conditions. As research continues to deepen our understanding of bio-
stimulants, their application can be further optimized to address climate
change and resource scarcity while supporting agricultural productivity.
5.1. Priming strategies with biostimulants against water and thermal
stresses
Priming plants with biostimulants has emerged as an effective
strategy to enhance their resilience to abiotic stresses. Seed priming is a
pre-sowing strategy to improve seed germination, early seedling estab-
lishment, and resilience to abiotic stress. Exposing seeds to controlled
stimuli before sowing triggers metabolic and physiological adjustments
that enhance water uptake, boost enzymatic activity, and strengthen
Table 1
A summary of the mechanisms through which seed priming biostimulants enhance plant molecular and physiological responses to heat and drought stress.
Type of Biostimulant Plant Species Application method and dose Biostimulatory Mechanisms Targeted
stress
References
Ascophyllum nodosum extract Spinacia oleracea Seed priming: 1.5–12 mL l
-1
solution
Improved chlorophyll content, antioxidant
activity, and overall plant growth
Drought (Pereira dos Anjos
Neto et al., 2020)
Azospirillum brasilense, Bacillus lentus,
and Pseudomonas sp. mixture
Ocimum basilicum Seed inoculation: 10⁸ CFU
mL
-1
Improved root growth, enhanced nutrient
absorption under reduced water stress
conditions
Drought (Heidari and
Golpayegani,
2012)
Azotobacter Triticum aestivum Seed inoculation: 10⁸ CFU
mL
-1
Enhanced chlorophyll retention, improved
WUE, increased antioxidant defenses and
resilience to drought and salinity
Drought,
salinity
(Chaudhary et al.,
2013)
Bacillus cereus Solanum
lycopersicum
Seed priming: bacterial
suspension, applied before
sowing
Activation of antioxidant pathways,
reduced lipid peroxidation, improved
growth
Heat,
drought
(Khan et al., 2020)
Bacillus polymyxa Lycopersicon
esculentum
Seed priming: 5 mL l
-1
solution (10
8
CFU mL
-1
)
Enhanced photosynthetic efciency,
reduced oxidative damage, improved
biomass
Heat (Shintu & Jayaram,
2015)
Commercial Bacillus strains
(B. amyloliquefaciens, B. licheniformis,
and Brevibacillus laterosporus)
Zea mays Soil application: 2 L ha
-1
during planting
Enhanced drought resilience, proline
biosynthesis, phenylpropanoid pathway
activation
Drought (Lephatsi et al.,
2022)
γ-Aminobutyric Acid (GABA) Helianthus annuus Seed priming: 2
μ
mol l
-1
for
24 h
Improved osmotic adjustment, reduced
oxidative stress, enhanced recovery
Heat,
drought
(Abdel Razik et al.,
2021)
γ-Aminobutyric Acid (GABA) Trifolium repens Seed priming: 2
μ
mol l
-1
for
24 h
Enhanced germination, root length,
antioxidant enzyme activity, and reduced
oxidative damage
Heat,
drought
(Zhou et al., 2021)
Chlamydomonas reinhardtii (cc124) and
Chlorella spp. (MACC-360, MACC-38)
Medicago
truncatula
Seed priming: 1 g l
-1
for 20
min; Root drenching: 0.05 g
l
-1
weekly for 35 days
Promotion of auxin-cytokinin homeostasis,
enhanced growth and resilience
Heat,
drought
(Gitau et al., 2021)
Enterobacter spp. Abelmoschus
esculentus
Seed treatment: 10⁵ CFU per
seed
Improved seed germination and early
development under stress
Drought (Roslan et al.,
2020)
Karrikins (KAR) Sapium sebiferum Seed priming: 1 nmol l
-1
solution for 24 h
Improved drought tolerance, antioxidant
defenses, delayed senescence
Drought (Shah et al., 2020)
Klebsiella aerogenes Triticum aestivum Seed treatment: 10⁵ CFU per
seed
Activated stress pathways, improved
proline levels, reduced ROS damage
Drought (Shafque et al.,
2023)
Rhizophagus irregularis AMF Lactuca sativa,
Solanum
lycopersicum
Soil inoculation: 10 g of
inoculum per plant
Improved nutrient uptake, upregulation of
ABA-responsive genes, reduced oxidative
damage
Drought (Ruiz-Lozano et al.,
2016)
Silicon (Si) Oryza sativa Seed priming: 7 mg l
-1
/ Soil
application: 300 kg ha
-1
Improved seed germination, enhanced
fruit yield, and water productivity
Drought (Chakma et al.,
2021)
Spirulina maxima extract Oryza sativa Seed priming: 2.5–5 g l
-1
aqueous extract
Enhanced germination, chlorophyll levels,
root elongation, and antioxidant activity
Heat,
salinity
(Sornchai et al.,
2014)
Spirulina platensis extract Triticum aestivum Seed priming: 100 mL l
-1
aqueous extract for 24 h
Enhanced photosynthesis, increased
photosynthetic pigments, improved WUE
Drought (Elnajar et al.,
2024)
Trichoderma harzianum Cuminum cyminum Root inoculation: 10 spores
mL
-1
Enhanced root biomass, improved nutrient
uptake, increased drought resilience
Drought (Piri et al., 2019)
Ulva lactuca extract Allium cepa Seed priming: 50 mL l
-1
aqueous extract
Enhanced chlorophyll levels, leaf area, and
root development under water stress
Drought (Muhie et al.,
2020)
Citrullus lanatus Seed priming: 30–80 mL l
-1
aqueous extract for 24 h
Improved germination rates, reduced
oxidative stress, enhanced root
development
Drought (Radwan et al.,
2023)
P. Carillo
Plant Stress 16 (2025) 100802
6
antioxidant defences (Paparella et al., 2015). As a result, primed seeds
show faster and more uniform germination, better-developed root and
shoot systems, and improved tolerance to drought and heat stress
(Savvides et al., 2016). Additionally, seed priming can induce stress
memory, allowing seedlings and young plants to enact a quicker and
more efcient defence response when exposed to environmental
stressors (Liu et al., 2022). The use of biostimulants in priming has
gained increasing attention also for due to their ability to further
enhance stress tolerance and plant performance through multiple
physiological and biochemical pathways The process integrates
biochemical, physiological, and molecular adaptations, ensuring plants
can efciently manage oxidative stress, osmotic imbalance, and
heat-induced cellular damage (Cocetta et al., 2022). Different types of
biostimulants offer distinct benets for priming (Table 1). All mea-
surement units have been standardised across different biostimulant
applications to ensure consistency and facilitate direct comparison of
treatment effects.
5.1.1. Seed priming with algal extracts
Algal extracts have demonstrated signicant effectiveness in pro-
moting seed germination and improving plant resilience to abiotic
stresses such as water scarcity, salinity and elevated temperatures. In
wheat, the grains priming with Spirulina platensis aqueous extract at a
100 ml l
-1
concentration signicantly enhanced plants drought toler-
ance. Treated plants exhibited up to a 53.4% increase in photosynthetic
efciency, enabling better carbon assimilation under water-limited
conditions. Additionally, the treatment improved plant vigour, with
drought-stressed cultivars showing a 153% rise in economic yield
compared to untreated controls. Growth parameters such as spike
number increased by 76.5%, while physiological traits like gas exchange
efciency and carbohydrate content improved signicantly (Elnajar
et al., 2024). Research shows that algal-based extracts enhance the
accumulation of osmolytes, including proline and GB, which are vital for
osmotic regulation during stressful conditions. Additionally, they
enhance the activity of key antioxidant enzymes, including ascorbate
peroxidase (APX), catalase (CAT), glutathione reductase (GR), peroxi-
dase (POD), and superoxide dismutase (SOD), helping to reduce the
oxidative damage in plant cells (Mughunth et al., 2024). This dual action
signicantly reduces oxidative stress, safeguarding cellular integrity
under adverse conditions (Carillo, Ciarmiello, et al., 2020). They also
increase photosynthetic pigments like chlorophylls and carotenoids and
promote larger leaf areas, longer shoots, and deeper root systems
(Parmar et al., 2023). In Medicago truncatula, treatments with
microalgae-based biostimulants, including Chlamydomonas reinhardtii
(cc124) and Chlorella spp. (MACC-360 and MACC-38), applied through
seed priming (1 g/L for 20 min) and weekly root drenching (0.05 g/L for
35 days), demonstrated strain-specic effects on auxin and cytokinin
homeostasis, signicantly inuencing plant growth and development.
Among the strains, Chlorella MACC-360 stood out for its ability to
enhance axillary shoot development, increase leaf size, and boost overall
biomass. This was attributed to its production of growth-promoting
phytohormones and exopolysaccharides (EPS), which improved
root-microbe interactions and enriched soil properties, further ampli-
fying its growth-promoting effects. In contrast, Chlorella MACC-38 and
C. reinhardtii exhibited more variable outcomes. While C. reinhardtii
enhanced micronutrient content in roots and leaves, it also induced
developmental irregularities, such as delayed owering and bifurcation,
likely due to disruptions in auxin-regulated gene expression (Gitau et al.,
2021). These ndings highlight the species or strain-specic ability of
(micro)algae to modulate plant growth dynamics, enhancing develop-
ment and improving plant resilience to environmental stresses. For
instance, Spirulina maxima extract (2.5–5 g L⁻¹) substantially enhanced
germination and early seedling development in Oryza sativa under
salinity and heat stress by enhancing chlorophyll levels, root elongation,
and antioxidant activity (Sornchai et al., 2014). In T. aestivum, seed
priming with Ulva lactuca extract (10–100 mL l
-1
10%) for 12 h boosted
drought and salinity resilience by raising proline levels, reinforcing
antioxidant systems, and promoting WUE as well as root development
(Ibrahim et al., 2014). Other studies demonstrated similar benets in
different crops. In Allium cepa, Ulva lactuca extract (5%) promoted the
germination of seeds and seedling vitality under water stress, leading to
increased chlorophyll levels, leaf area and root development (Muhie
et al., 2020). A study evaluating the effect of ve Ascophyllum nodosum
extract concentrations (0.0, 0.15, 0.30, 0.60, and 1.2%) on Spinacia
oleracea seeds found that 0.3% extract priming determined the highest
germination percentage (+18%), faster germination speed (+22%), and
improved seedling vigor (+20%) compared to non-primed seeds under
heat stress (30 ◦C). Additionally, the priming reduced H₂O₂ and MDA
levels by 25% and 30%, respectively, under heat stress, indicating
enhanced stress tolerance mechanisms (Pereira dos Anjos Neto et al.,
2020). An aqueous extract of Ulva lactuca (3–8%) applied to Citrullus
lanatus (Gurum) improved germination rates and seedling vigor under
drought conditions by enhancing proline content and antioxidant
enzyme activity, reducing oxidative damage (Radwan et al., 2023). Xu
et al. (2018) discovered a CGTCA motif located in the promoter region of
choline monooxygenase (CMO) and betaine aldehyde dehydrogenase
(BADH) genes in Citrullus lanatus, which is responsive to methyl jasm-
onate (MeJA). This nding demonstrated that C. lanatus cells treated
with MeJA could synthesize GB even without environmental stress
(Annunziata et al., 2019). The activation of this MeJA-GB synthesis
cascade may offer a valuable perspective on the potential modes of ac-
tion through which algal biostimulants may enhance plant stress toler-
ance. This mechanism also highlights the role of GB as a particularly
intriguing compound in stress adaptation. It is notably synthesised and
accumulated in halophytes and some cultivated plants like Poaceae. Its
amphoteric nature allows it to remain electrically neutral across a wide
physiological pH range and is highly water-soluble. Under stress con-
ditions, GB acts as an osmoregulator, stabilizes macromolecules and
their functions, maintains membrane integrity, and mitigate cellular
damage, playing many essential roles (Annunziata et al., 2019). GB
typically accumulates in young tissues during prolonged stress. How-
ever, plants synthesize it sparingly and only when necessary since it is
non-recyclable and gradually diluted (Carillo et al., 2008). A delay in GB
synthesis under drought and/or salinity can lead to structural disrup-
tions, particularly at the root apex, with signicant vacuolization,
disorganized apical tissue, and slight plasmolysis due to reduced cell
adhesion. These changes can inhibit growth and differentiation, severely
affecting plant development (Annunziata et al., 2017; Duan et al., 2010).
The supply of GB (5 mM) to the culture medium has been shown to
effectively counteract osmotic stress in barley roots by preserving root
apex structure and mitigating growth inhibition, as demonstrated by
Cuin and Shabala (2005). Similarly, in Citrullus lanatus, the activation of
the MeJA-GB pathway by algal extracts likely prevents such structural
damage by promoting early GB synthesis and maintaining cellular
integrity under stress conditions (Radwan et al., 2023; Xu et al., 2018).
5.1.2. Seed priming with microbial biostimulants
Microbial biostimulants, including arbuscular mycorrhizal fungi
(AMF) and plant growth-promoting bacteria (PGPB), boost nutrient
acquisition through processes such as nitrogen xation and the solubi-
lization of phosphorus. These microorganisms form mutually benecial
relationships with plants, establishing dense hyphal networks that in-
crease root nutrient absorption. They also reinforce cell walls and create
protective biolms on root surfaces, improving water retention and
salinity resilience under thermal and water stress conditions (Fusco
et al., 2022). These interactions activate crucial physiological and
metabolic pathways, promoting the accumulation of protective com-
pounds like proline, AA, and GSH while boosting the activity of anti-
oxidant enzymes (SOD, POD, CAT). Microbial biostimulants also
enhance the uptake of essential major nutrients (N, P, K, Ca, Mg) and
trace elements (Fe, Zn, Cu, Mn, B), improving chlorophyll synthesis,
photosynthetic efciency, carbon assimilation, and biomass production
P. Carillo
Plant Stress 16 (2025) 100802
7
(Papa et al., 2022). This supports root development, leaf expansion, and
overall plant growth, ultimately increasing resilience to environmental
stresses like drought and heat, ensuring improved survival and pro-
ductivity under challenging conditions (Fadiji et al., 2021). For example,
Rhizophagus irregularis (strain EEZ 58) enhanced drought tolerance in
Lactuca sativa, cv. Romanain and S. lycopersicum cv. Reimlams Rhums
through arbuscular mycorrhizal (AM) symbiosis . At sowing, soil was
supplied with 10 grams of inoculum or autoclaved inoculum, respec-
tively, eachcontaining approximately 60 infective propagules per gram,
ensuring the establishment of the symbiosis. Under drought stress, root
colonization increased signicantly, reaching 63% in lettuce and 54% in
tomato after 8 weeks. This enhanced colonization improved shoot dry
weight (+26% in lettuce and +17% higher in tomato) improved
photosystem II efciency and stomatal conductance, supporting photo-
synthetic activity under drought. AMF treated plants also showed
increased ABA levels, upregulation of ABA biosynthesis genes such as
LsNCED2, and activation of drought-responsive genes, including dehy-
drins (DHNs) from the late embryogenesis abundant (LEA) family,
which stabilized membranes mitigated oxidative stress damage. Simi-
larly, strigolactones (SLs), essential for root-fungus communication,
were signicantly upregulated in mycorrhized plants under drought.
Tomato plants in symbiosis with R. irregularis showed higher levels of
SLs like solanacol and didehydro-orobanchol, and upregulation of the SL
biosynthesis gene SlCCD7. By stimulating the ABA and strigolactone
signalling pathways and enhancing critical stress-response mechanisms,
AM symbiosis emerges as an effective and sustainable approach to boost
crop resilience in water-limited conditions (Ruiz-Lozano et al., 2016).
Klebsiella aerogenes SH-8 (10
5
CFU/Petri) signicantly improved
drought tolerance in T. aestivum by enhancing proline accumulation,
boosting antioxidant enzyme activity, and increasing nutrient uptake,
resulting in better root development and overall plant growth under
drought conditions (Shafque et al., 2023). Similarly, Enterobacter spp.
(10
4
CFU seed
-1
) enhanced seed germination and initial seedling
development in Abelmoschus esculentus by increasing chlorophyll con-
tent and leaf area, thereby supporting plant vigor during water decit
(Roslan et al., 2020). In Cuminum cyminum, a combination of Pseudo-
monas uorescence (10
8
CFU mL
-1
) and Trichoderma harzianum (10
spores mL
-1
) strains effectively enhanced root biomass and improved
nutrient uptake, signicantly increasing the plant’s drought resilience
(Piri et al., 2019). In Hordeum vulgare, the application of Hartmannibacter
diazotrophicus (10⁸ CFU mL⁻¹ seed⁻¹) resulted in improved chlorophyll
content, greater root development, and enhanced antioxidant activity,
supporting better growth under water scarcity (Suarez et al., 2015).
Bacillus polymyxa (5 mL l
-1
of a solution of 10
8
CFU mL
-1
) enhanced the
stress tolerance of Lycopersicon esculentum through seed priming. Primed
seeds were sown in garden pots, and after 21 days of vegetative growth,
plants underwent a 3-day water stress treatment, followed by regular
irrigation This treatment led to higher proline levels, improved photo-
synthetic efciency, and increaed antioxidant defences, which mitigated
the effects of heat stress (Shintu, 2015). Additionally, Bacillus cereus
priming improved the tolerance to elevated temperatures in
S. lycopersicum by enhancing the transcription of SlWRKY33b and
autophagy-related genes (SlATG5), and upregulating genes encoding
heat-responsive TFs (SlHsfA1a) and potassium transporters (SlHKT1).
This mechanism supported root development and reduced oxidative
damage under elevated temperatures (Khan et al., 2020). The inocula-
tion of Ocimum basilicum seeds with a bacterial suspension of Azospir-
illum brasilense, Bacillus lentus, and Pseudomonas sp. (10
8
CFU mL
-1
)
mixed with perlite enhanced drought tolerance by promoting shoot and
root growth and supporting nutrient absorption, even under reduced
water availability (60% and 40% of the FC vs 80% control) (Heidari and
Golpayegani, 2012). Additionally, inoculation of T. aestivum seeds with
Azotobacter (10
8
CFU mL
-1
) improved WUE and chlorophyll retention,
boosting plant resilience to drought and salinity (Chaudhary et al.,
2013). A commercial biostimulant formulation, consisting of ve Ba-
cillus strains (one strain of Bacillus amyloliquefaciens, two strains of
Bacillus licheniformis, and two strains of Brevibacillus laterosporus) was
supplied at a dose of 2 L per hectare, delivering directly into the soil
ensuring direct contact with the seed of Zea mays at planting. The con-
sortia signicantly enhanced drought resilience by activating
stress-responsive pathways. This priming led to upregulating ZmDREB2
and ZmP5CS, essential for osmotic adjustment via proline biosynthesis.
The consequently elevated proline levels, coupled with a 6.3-fold in-
crease of the expression of phenylalanine ammonia-lyase (PAL), and a
0.6-fold decrease of avone synthase (FSNII), further enhanced drought
tolerance. PAL acts as a pivotal enzyme in linking primary and sec-
ondary metabolism through the phenylpropanoid pathway, which con-
verts phenylalanine into cinnamic acid, a fundamental precursor for
important phenylpropanoids. Conversely, FSNII facilitates the direct
conversion of avanones into avones, contributing to the production of
secondary metabolites that support growth and defence mechanisms
under water stress (Lephatsi et al., 2022). In Z. mays, Bacillus sp. MGW9
(10
8
CFU mL
-1
) demonstrated similar benets by supporting root system
development and antioxidant activity, improving resilience to water
stress (Li et al., 2021).
5.1.3. Seed priming with plant extracts, protein hydrolysates and
metabolites
Seed priming with plant extracts, protein hydrolysates (PHs), or
metabolites is an established approach to enhance germination and
improve water stress tolerance. This technique activates key physio-
logical mechanisms, such as DNA repair, cell division, and antioxidant
enzyme activity, fostering uniform germination, robust shoot develop-
ment, and improved relative water content (RWC). By reducing ROS and
MDA accumulation, it mitigates oxidative stress, stabilizes cellular
membranes, and enhances drought resilience, enabling faster recovery
after rehydration (Ahmed et al., 2019; Wang et al., 2022). For example,
γ-aminobutyric acid (GABA) pretreatment at 2
μ
mol l
-1
signicantly
enhanced drought tolerance in Trifolium repens (white clover) seeds,
increasing germination percentage by over 20% and root length by 15%
under water-limited conditions. This treatment also enhanced soluble
sugars and antioxidant metabolites’ levels (e.g., AA and GSH) while
enhancing antioxidant enzyme activities (e.g., CAT, POD, and SOD) by
25–40%. Furthermore, GABA priming increased DHN proteins and
DREB TFs levels, improving osmotic adjustment and reducing oxidative
damage during germination (Zhou et al., 2021). Similarly, O. sativa
seeds exposed to MeJA under osmotic stress induced by polyethylene
glycol exhibited increased metabolic regulation and resilience, high-
lighting the effectiveness of priming strategies in alleviating abiotic
stress (Sheteiwy et al., 2018). These ndings underscore the effective-
ness of seed priming in enhancing stress tolerance through metabolic
and antioxidant modulation. Also priming seeds of O. sativa with 0.5
mmol l
-1
GABA markedly enhanced germination and seedling develop-
ment under osmotic stress caused by polyethylene glycol (30 g l
-1
PEG
6000). GABA priming enhanced water relations and photosynthetic ef-
ciency while increasing the levels of GABA, soluble sugars, starch and
proteins. Additionally, antioxidant enzyme activities (e.g., GR) and
detoxication-related enzymes were signicantly enhanced, thus
reducing oxidative stress. Notably, GABA-treated seeds exhibited lower
proline levels and reduced free radicals and MDA under stress.
Furthermore, the transcription of stress-responsive genes, particularly
OsCIPK02, OsCIPK07, and OsCIPK09, was upregulated under osmotic
stress, highlighting the molecular basis of GABA-mediated stress toler-
ance. The OsCIPK (Calcineurin B-like Protein-Interacting Protein Ki-
nases) are genes in rice that respond to stress and function in calcium
signalling pathways. These kinases interact with Calcineurin B-like
proteins (CBLs) to regulate ion homeostasis, and to activate antioxidant
defences under osmotic stress conditions. They enable cellular adapta-
tion and improve plant resilience during water-decit and salinity stress
(Xiang et al., 2007). Karrikins (KARs), active components of smoke, have
been recognized as valuable tools for improving tolerance to abiotic
stresses. KAR1 at a concentration of 1 nM signicantly improved
P. Carillo
Plant Stress 16 (2025) 100802
8
germination of Sapium sebiferum seeds and seedling vitality under os-
motic stress (100–300 mmol l
-1
mannitol) conditions. The primed seeds
showed higher activities of antioxidative enzymes including SOD, POX,
CAT, and APX, lower levels of H
2
O
2
, MDA, and electrolyte leakage (EL),
together with improved germination rates (Shah et al., 2020).
Silicon-based treatments, including seed priming and soil application,
were evaluated for their effectiveness in improving drought resilience in
grape tomato (S. lycopersicon L. var. cerasiforme). Monosilicic acid
(MSA) proved ineffective under severe drought conditions (50% FC)
irrespective of the application method or dose. However, MSA applied as
seed priming at a concentration of 0.25 mM or applied to the soil at a
rate of 300 kg/ha boosted fruit yield and irrigation water productivity
(IWP) under moderate (75% FC) and optimal (100% FC) moisture levels
(Chakma et al., 2021).
In Pisum sativum, seed priming with 50 mg L⁻¹ SA improved seedling
vigor and physiological stability under heat stress conditions (43 ◦C).
SA-treated seeds exhibited higher germination rates, improved seedling
elongation, higher fresh and dry biomass, and increased relative water
content. At the physiological level, SA enhanced chlorophyll retention
and membrane stability, supporting photosynthetic efciency and water
balance (Tamindˇ
zi´
c et al., 2023). In Hordeum vulgare cv. Haider-93, seed
priming with 20 mM GB signicantly improved germination, shoot
biomass, and water retention under high temperatures. Seedlings from
GB-treated seeds exhibited higher shoot dry weight and net photosyn-
thesis (Pn) while maintaining lower membrane permeability than un-
treated plants. Additionally, GB reduced ion leakage, particularly Ca²⁺,
Table 2
An overview of the mechanisms by which foliar spray and root drenching biostimulants enhance plant molecular and physiological adaptations to heat and drought
stress. All concentrations have been standardized to ensure comparability across treatments.
Type of biostimulant Plant species Application method and dose Biostimulatory mechanisms Targeted
stress
References
Ascophyllum nodosum
extract
Arabidopsis thaliana Foliar spraying: 2 mL l
-1
on days 25 and 27
after germination
Enhanced RWC, reduced oxidative
stress, regulated ROS, improved cell
cycle activity
Drought (Rasul et al., 2021)
Ascophyllum nodosum
and Laminaria digitata
extracts
Solanum
lycopersicum
Foliar spraying: 0, 2.5, and 5 mL L⁻¹ during
critical growth stages
Improved fresh/dry weight, enhanced
RWC, antioxidative enzyme activity,
stabilized chlorophyll
Drought (Campobenedetto
et al., 2021)
Ascophyllum nodosum
extract and Silicon (Si)
Solanum
lycopersicum
Foliar: 0, 1.25, 2.5, 3.75, 5 mL L⁻¹; Silicon: 60
kg ha⁻¹ applied to soil
Improved water productivity, reduced
ion leakage, stabilized biomass,
enhanced soluble solids and fruit
quality
Drought (Ahmed et al., 2023)
Bacillus amyloliquefaciens
SN13
Oryza sativa Root inoculation: 10 ml l
-1
solution Increased osmoprotectants, reduced
oxidative damage, upregulated stress-
responsive genes
Heat,
drought
(Tiwari et al., 2017)
Bacillus mycoides strain
A3
Arabidopsis thaliana Foliar spraying: 1 ×10⁸ CFU mL
-1
applied
twice
Improved root architecture, increased
antioxidant enzyme activity, reduced
EL and MDA
Heat,
drought
(Kurniawan and
Chuang, 2022)
Bacillus zanthoxyli HS1 Brassica rapa spp.
pekinensis, Cucumis
sativus
Soil inoculation: 1.3 ×10⁷ CFU mL
-1
or VOC
exposure
Enhanced shoot/root growth, improved
antioxidant activity, modulated
stomatal conductance, reduced stress
sensitivity
Heat,
drought
(Barghi and Jung,
2024)
β-Aminobutyric Acid
(BABA)
Linum usitatissimum
(linseed)
Seed priming: 50–100
μ
mol l
-1
; Foliar
spraying: not specied
Enhanced CAT/POX activity, increased
proline levels, reduced oxidative
damage, improved yield
Drought (Yasir et al., 2023)
γ-Aminobutyric Acid
(GABA)
Helianthus annuus Foliar spraying: 1 mmol l
-1
applied 20 days
after treatment
Improved gas exchange, enhanced PSII
efciency, reduced oxidative stress,
increased seed yield
Heat,
drought
(Abdel Razik et al.,
2021)
Glutamate Brassica napus Root drenching: daily application, 2 mL of 20
mmol l
-1
glutamate
Enhanced proline levels, improved
RWC, boosted drought resilience
Drought (La et al., 2020)
Karrikins (KAR1) Agrostis stolonifera Foliar spraying: 100 nmol l
-1
Increased RWC, delayed senescence,
enhanced antioxidant enzyme activity
Drought (Tan et al., 2023)
Mannitol Triticum aestivum Root drenching: 100 mmol l
-1
applied for 24
h
Enhanced antioxidant enzyme activity,
improved tolerance to osmotic and
salinity stress
Drought,
salinity
(Seckin et al., 2009)
Nannochloris sp.
biostimulant
Lycopersicon
esculentum
Foliar spraying: 0.5% solution (0.5 kg 100 l
-1
ha
-1
) applied twice (2nd and 29th days post-
repotting)
Enhanced root/leaf development,
reduced ROS, improved growth under
drought stress
Drought (Oancea et al., 2013)
Protein hydrolysates
(PHs)
Capsicum annuum Fertigation: 1.5 mL L⁻¹ Maintained chlorophyll content,
improved biomass, enhanced
antioxidant defences
Drought (Agliassa et al.,
2021)
Satureja hortensis Foliar spraying: 2 mL l
-1
applied twice Increased proline, improved pigment
levels, enhanced yield, supported
growth under drought
Drought (Rezaei-Chiyaneh
et al., 2023)
Solanum
lycopersicum
Foliar spraying: 3 mL L⁻¹ weekly Activated defence mechanisms,
improved WUE, reduced stress-related
agents
Drought (Leporino et al.,
2024)
Septoglomus constrictum
AM fungi
Solanum
lycopersicum
Root inoculation: 30 g inoculum/pot (27–32
spores g
-1
)
Improved biomass, enhanced
antioxidant enzyme activity, reduced
MDA, upregulated stress-responsive
genes
Heat,
drought
(Duc et al., 2018)
Thuricin 17 (Th17) Brassica napus Root drenching: 1 nmol l
-1
(severe drought)
or 10 pmol l
-1
(heat +drought)
Enhanced root length, increased CAT/
POD/SOD activity, reduced oxidative
stress, improved water uptake
Heat,
drought
(Nazari and Smith,
2023)
Urea, zinc, iron, silicon Triticum aestivum Seed priming: urea (20 g l
-1
), FeSO₄ (50 mg l
-
1
), ZnSO₄ (50 mg l
-1
), silicon (20 mg l
-1
);
Foliar spray: urea (40 g l
-1
), silicon (112 mg l
-
1
), FeSO₄ (6 g l
-1
), ZnSO₄ (4 g l
-1
)
Increased grain yield, improved
nutrient biofortication (Zn and Fe),
reduced oxidative damage
Drought (Moradi and
Siosemardeh, 2023)
P. Carillo
Plant Stress 16 (2025) 100802
9
K⁺, and NO₃⁻, helping to stabilise cell membranes under heat stress.
These ndings indicate that GB priming enhances seedling vigour by
preserving water balance, photosynthetic efciency, and cellular sta-
bility, potentially improving crop establishment in high-temperature
environments (Wahid and Shabbir, 2005). Pre-sowing seed treatment
with a commercial biostimulant based on lignin derivatives, amino
acids, and molybdenum, improved germination and early growth in
Cucumis sativus L., particularly under heat stress (35 ◦C). After 48 h,
treated seeds showed higher germination (+6.54%) and fresh biomass
(+13%). The treatment upregulated the expression of antioxidant
enzyme genes (CuZnSOD +1.78, MnSOD +1.75, CAT +3.39) and
increased non-protein thiols (+20%), reduced oxidative damage
(decrease of H₂O₂ levels by 70% at 28 ◦C and 80% at 35 ◦C) and stim-
ulated metabolic activity (37% increase of isocitrate lyase activity), thus
improving seedling establishment under heat stress (Campobenedetto
et al., 2020).
5.2. Foliar spraying and root drenching for managing water and thermal
stresses
Foliar spraying and root drenching, which involve applying bio-
stimulant solutions directly to plant leaves or roots, have proven effec-
tive in alleviating heat and drought stress. Such methods improve plant
physiological and biochemical functions by enhancing WUE, increasing
nutrient absorption, and triggering stress-response pathways. Under
elevated temperatures, foliar applications can boost cellular defences by
increasing HSPs and antioxidants, whereas root drenching supports root
health and water uptake. Under drought stress, these approaches
improve stomatal control and increase the accumulation of compatible
osmolytes such as proline and GB, thereby stabilising cell structures
during water scarcity and minimising oxidative damage. Delivering
active substances directly to key plant tissues through foliar and root
applications mitigates immediate stress impacts and enhances long-term
resilience, promoting sustainable yields even under adverse climatic
conditions (Table 2).
5.2.1. Algal extracts for foliar and root applications
Algal extracts and biostimulants derived from microalgae can
enhance plant resilience against drought and heat stress when applied
via foliar spraying or root drenching. These biostimulants demonstrate a
range of benecial effects, including an increase of the osmoprotectants
proline and GB, stimulation of the activity of antioxidant enzymes (e.g.,
APX, CAT, GR, POD and SOD), and promotion of photosynthetic
pigment synthesis (chlorophylls and carotenoids). These biochemical
changes enhance photosynthesis, light protection, and water retention
while supporting growth metrics like leaf expansion, stem elongation,
and root growth. Together, these effects facilitate nutrient and water
uptake in stress-prone conditions (Carillo, Ciarmiello, et al., 2020).
An extract from Ascophyllum nodosum, foliar sprayed at a 2 mL l
-1
concentration on days 25 and 27 after germination, effectively allevi-
ated drought stress in A. thaliana, induced by withholding irrigation
between days 31 and 42 post-germination. Plants treated with the
extract exhibited reduced oxidative stress, as evidenced by lower ROS
accumulation, alongside improved water retention and stomatal regu-
lation compared to untreated controls. Enhanced RWC and decreased
ion leakage in leaves indicated greater cellular stability and reduced
damage. The presence of specic components like laminarin, fucoidan,
alginate, and ulvan in algal extracts enhances water stress tolerance
through the priming of antioxidant defences and regulation of ROS
levels. These compounds mitigate oxidative damage, stabilize cellular
structures, and promote photosynthetic efciency under stress condi-
tions. Molecular analyses also revealed that the extract facilitated cell
cycle activity by sustaining the expression of genes such as HISTONE H4
and Cyclin P2;1, a key regulator of cell division at the shoot apical
meristem (SAM) under drought conditions. Additionally, the extract
downregulated RD26, a stress-responsive growth suppressor, in the
SAM, ensuring its functionality and promoting sustained growth (Rasul
et al., 2021).
In S. lycopersicum, foliar applications of seaweed extracts derived
from A. nodosum and Laminaria digitata under water stress conditions
demonstrated signicant growth, yield, and stress tolerance improve-
ments. The seaweed extract was applied at concentrations of 0, 2.5, and
5 mL L⁻¹ during critical growth stages, with evaluations performed
under both optimal irrigation and drought conditions. The biostimulant-
treated plants showed increased fresh weight and dry matter content
and improved stem water potential, suggesting better water retention
and transport mechanisms. Biochemical adjustments included enhanced
activity of antioxidative enzymes (e.g., SOD and CAT), increased proline
levels, and stabilised chlorophyll concentrations. The treatments
improved RWC, mitigated oxidative damage, and sustained photosyn-
thetic efciency, ultimately enhancing fruit yield under drought stress
(Campobenedetto et al., 2021). Moreover, always in S. lycopersicum, the
simultaneous use of A. nodosum extract and silicon (Si) across different
soil moisture conditions (50%, 75%, and 100% FC) demonstrated
additional benets in mitigating water stress. The algal extract was used
as a foliar spray at concentrations of 0, 1.25, 2.5, 3.75, and 5 mL L⁻¹,
whereas silicon, as mono silicic acid, was applied to the soil at a dosage
of 60 kg ha⁻¹. This combination improved water productivity and fruit
quality traits, including higher soluble solids, rmness, and pH. Under
severe drought stress (50% FC), the treatments reduced yield losses,
oxidative damage, and ion leakage, while maintaining higher RWC and
promoting shoot and root biomass. Algal and Si applications resulted in
comparable yields at moderate (75%) and sufcient (100%) soil mois-
ture levels, highlighting their potential to sustain productivity across
varying water availability conditions (Ahmed et al., 2023).
In a greenhouse study, a microalgae-based biostimulant derived from
Nannochloris sp. 424–1 was tested on tomato plants (Lycopersicon escu-
lentum cv. Cristal F1) under water-stressed and non-stressed conditions.
The experiment included six treatments, with foliar sprays applied on
the 2nd and 29th days post-repotting. Plants were irrigated either every
ve days (non-stressed) or every two weeks (stressed) at 100% eld
capacity. The biostimulant, composed on a dry weight basis of proteins
(88.26%), proline (5.63%), carbohydrates (9.41%), and betaines
(0.023%), was applied as a 5 g l
-1
solution (0.5 kg 100 l
-1
ha
-1
) in
comparison with a commercial A. nodosum based extract. After eight
weeks, the Nannochloris sp.-based biostimulant effectively alleviated
water stress effects, promoting growth and productivity in a way com-
parable to or exceeding the commercial product. When tested on non-
water-stressed tomato plants, the microalgae-based biostimulant out-
performed the commercial seaweed extract in stimulating root length
(+2%) and leaf number (+13). Under water stress conditions, untreated
tomato plants experienced reductions of nearly 20% in height and over
25% in root length. However, the application of the microalgae-based
biostimulant mitigated these effects, reducing the negative impact on
height by nearly 50% and supporting root development. The microalgae
biostimulant and the commercial macroalgae product demonstrated
similar results during the 30-day owering and fruiting cycle (Oancea
et al., 2013). Therefore algal and microalgae-derived biostimulants
represent a sustainable and effective strategy to mitigate the effects of
abiotic stresses such as drought and heat. These biostimulants enhance
plant resilience by activating specic physiological and molecular
pathways, providing practical solutions to maintain agricultural pro-
ductivity in increasingly challenging environmental conditions.
5.2.2. Root inoculation with microbial biostimulants
Microbial biostimulants, including arbuscular mycorrhizal (AM)
fungi, may also improve plant tolerance to drought and heat stresses
through diverse biochemical and physiological mechanisms. In a
controlled experiment on S. lycopersicum, inoculation with Septoglomus
deserticola and Septoglomus constrictum at transplant was assessed for its
ability to alleviate combined drought (50% eld capacity) and heat
stress (42 ◦C for 6 h). AM inoculation was achieved by applying 30
P. Carillo
Plant Stress 16 (2025) 100802
10
grams of inoculum per pot containing 27–32 spores per gram.
S. constrictum signicantly improved plant biomass, increasing root and
shoot dry weight by 14% and 18% under drought and by 22% and 24%
under combined stress, respectively, compared to non-inoculated con-
trols. The maximum quantum efciency of photosystem II (Fv/Fm)
improved by 9% under combined stress, while stomatal conductance
and relative water content also increased. Leaf water potential showed a
21% improvement under drought, reecting better water retention.
Oxidative stress markers, including MDA and H
2
O
2
, were notably
reduced in AM plants: in particular, under combined stress,
S. constrictum inoculation lowered MDA by 18% and H
2
O
2
by 37%.
Enhanced activities of the antioxidant enzymes CAT, SOD, and POD
contributed to reduced oxidative damage; CAT activity, for instance,
increased by 87% in roots under drought and by 28% under combined
stress. At the molecular level, S. constrictum inuenced the expression of
stress-responsive genes. SlLOXD, involved in MeJA biosynthesis, was
upregulated by 32% under combined stress, supporting stress mitigation
pathways. Conversely, SlNCED, related to ABA synthesis, was down-
regulated, correlating with improved WUE (Duc et al., 2018) . Beyond
AM fungi, also plant growth promoting rhizobacteria (PGPR) may
modulate key biochemical and molecular pathways to increase plant
resilience to enviromental stresses. In particular, inoculation with Ba-
cillus amyloliquefaciens SN13 (SN13) at 1% signicantly enhanced the
resilience of O. sativa seedlings to combined heat and drought stress.
SN13-treated plants demonstrated higher levels of osmoprotectants,
reduced oxidative damage, and enhanced stress-responsive gene
expression, collectively improving their ability to withstand stress.
Proline levels increased by 20% in SN13-inoculated seedlings compared
to untreated controls, especially within the rst 3 h of stress exposure.
Total soluble sugars also showed a remarkable rise, with levels
increasing by 328% at 3 h and maintaining 200–400% higher concen-
trations at 24 h. SOD and CAT activities were signicantly enhanced,
protecting cells from ROS, and reducing oxidative damage as also evi-
denced by the lower MDA levels, indicating reduced lipid peroxidation
and improved membrane integrity. On a molecular level, SN13 activated
key stress-responsive genes, including those encoding for DHN and LEA
proteins, and glutathione S-transferases (GSTs), which were upregulated
by 7–12-fold during stress. These genes played a crucial role in stabi-
lizing cellular structures, mitigate oxidative damage, and maintaining
cellular functions under adverse conditions (Tiwari et al., 2017).
Notably, GSTs play a crucial role in plant defence mechanisms by
detoxifying ROS through conjugation with glutathione. This process
protects cellular structures, maintains redox balance, and supports stress
signalling, thereby enhancing plant resilience under abiotic stress con-
ditions (Edwards et al., 2000). Two-week-old soil-grown seedlings of
A. thaliana Col-0 were treated twice with Bacillus mycoides strain A3
(BmA3) at 10
8
CFU mL
-1
before withholding water for 7 days or
exposing them to 45 ◦C for 20 min, followed by recovery at 23 ◦C for 24
h. BmA3-treated plants under water or thermal stresses showed lower
H₂O₂, MDA, and EL, and greater fresh weights and higher survival rates
after rewatering or 7 days post-recovery from thermal stress. BmA3
enhanced root architecture, increasing lateral roots and root hairs, and
maintained high chlorophyll levels thus boosting photosynthetic ef-
ciency and starch accumulation. Antioxidant defences were strength-
ened via higher activities of CAT, GPX, APX, and PAL, while phenolics,
avonoids, and glucosinolates increased signicantly. Stress-responsive
genes, including DREB2A and HsFA2, were upregulated alongside those
involved in antioxidant enzyme (APX, AOX), JA (MYC2, LOX1), and SA
(SARD1, CBP60 G) pathways, thereby improving adaptation and re-
covery under abiotic stress (Kurniawan and Chuang, 2022).
Root drenching with Bacillus cereus, applied to S. lycopersicum one
week after transplanting and repeated twice at ve-day intervals,
enhanced heat tolerance by stimulating the transcription of SlWRKY33b
and autophagy-related genes (SlATG5), as well as upregulating heat-
responsive transcription factors (SlHsfA1a) and potassium transporters
(SlHKT1). The activation of these pathways supported root development
and mitigated oxidative damage when plants were later exposed to
elevated temperatures (Khan et al., 2020). The application of Bacillus
zanthoxyli HS1, a strain isolated from cucumber rhizosphere soils, and its
volatile organic compounds (VOCs), enhanced heat stress tolerance in
Brassica rapa spp. pekinensis cv. Ryeong-gwang and Cucumis sativus cv.
Jo-eun-baeg-log-da-da-gi. Seedlings were transplanted into soil pre-
treated with B. zanthoxyli HS1 at 1.3 ×10
7
CFU mL-1 or exposed to its
VOCs before being subjected to 36 ±2 ◦C for two days, followed by
recovery. Bacillus HS1 treatment improved shoot and root growth,
increased chlorophyll content, and reduced proline levels, indicating
reduced stress sensitivity and more efcient stress response. The efcacy
of HS1 was also compared with azelaic acid (AzA), a well-known primer
for systemic acquired resistance (SRA). HS1 proved more effective in
mitigating abiotic stress, while AzA disrupted HS1
′
s protective effects
when co-applied. HS1 also upregulated the expression of key antioxi-
dant genes (APX, CAT, and SOD), promoted callose deposition thus
improving structural defence, modulated stomatal conductance for
water conservation, and reinforced systemic tolerance through
VOC-mediated signalling (Barghi and Jung, 2024). Some microbes, such
as Pseudomonas spp. and Achromobacter piechaudii, identied in
semi-arid regions, have demonstrated remarkable potential in confer-
ring drought and heat tolerance to different crops like T. aestivum
(Shakir et al., 2012), S. lycopersicum (Chen et al., 2014), and Capsicum
annuum (Maxton et al., 2018), Vitis vinifera (Duan et al., 2021) These
PGPR reduce ethylene levels in plants by hydrolyzing its precursor ACC
into ammonia and
α
-ketobutyrate through the action of ACC deaminase
(ACCD). This enzymatic process not only mitigates the accumulation of
stress-induced ethylene, which can inhibit plant growth and lead to
premature senescence and abscission, but also enables PGPR to utilize
ACC as a carbon and nitrogen source. By reducing ethylene synthesis,
these PGPR improve RWC, photosynthetic pigment concentrations, and
leaf area while lowering proline levels. The reduction in proline reects
alleviated stress and enhanced physiological adaptation. The dual role of
microbial ACCD in reducing the synthesis of ethylene and using its
precursor as a nutrient source makes it a cost-effective and sustainable
strategy for improving plant resilience under abiotic stress. Moreover,
the inducibility of the bacterial AcdS gene, which encodes ACCD, allows
for adaptive regulation under diverse environmental conditions, making
it a valuable tool for enhancing crop performance in stress-prone envi-
ronments. This synergy of microbial and genetic strategies underscores
the pivotal role of ACCD in fostering resilient agricultural systems (Singh
et al., 2022). These ndings align with Induced Systemic Tolerance
(IST), where PGPR triggers physical and biochemical adjustments to
enhance resilience against abiotic stresses. Accordingly, in Sorghum
bicolor Moench, the priming with 77 non-pathogenic rhizobacterial
isolates, applied three weeks after planting as 1 mL of a 10
8
CFU mL
−1
of
cell suspension per plant, resulted in signicant metabolomic reprog-
ramming (Carlson et al., 2020). UHPLC
–
HDMS analysis identied key
biomarkers linked to drought stress tolerance, including enhanced
antioxidant capacity, modied root architecture driven by upregulated
synthesis of GA, ABA, and CK, and the early activation of stress signal-
ling pathways involving brassinolides, SA, JA, sphingosine, and psy-
chosine (two lipid molecules involved in cellular signalling and stress
responses). Moreover, osmolytes and epicuticular wax (docosanoic acid)
further contributed to drought tolerance. ACC deaminase activity
reduced ethylene levels, mitigating stress-induced growth inhibition
(Carlson et al., 2020).
5.2.3. Foliar and root treatments with metabolites and protein hydrolysates
Foliar treatments with 50 µmol L⁻¹ of IAA, GA, or BR, and 10 µmol L⁻¹
of CK were applied to Oryza sativa cultivars Fedearroz 67 (F67) and
Fedearroz 2000 (F2000) twice, 5 days before and 5 days after the onset
of heat stress, using 20 mL per plant with 1 mL l
-1
agricultural adjuvant
to ensure leaf coverage. Heat stress (40 ◦C day/30 ◦C night) was imposed
for 8 consecutive days in growth chambers, while control plants were
maintained under optimal conditions (30 ◦C day/25 ◦C night). CK
P. Carillo
Plant Stress 16 (2025) 100802
11
treatments signicantly enhanced chlorophyll content in heat-stressed
plants F67 (+38%) and F2000 (+43%) compared to untreated stressed
plants, while stomatal conductance in F2000 increased by 232%, indi-
cating better WUE. BR and CK applications reduced canopy tempera-
tures by 2–3 ◦C and oxidative damage, as evidenced by lower MDA
levels. Relative tolerance indices showed that CK (98%) and BR (61%)
were highly effective in maintaining water content, chlorophyll levels,
and overall plant vitality, demonstrating their potential as agronomic
strategies to mitigate heat stress in the two rice cultivars
(Pantoja-Benavides et al., 2021). Also, SA enhanced heat resilience in
grapevine (Vitis vinifera) by stabilising photosynthetic efciency and
accelerating post-stress recovery. Seedlings were foliar-sprayed with SA
0.1 mmol l
-1
24 h before heat stress and exposed to 43 ◦C for 5 h, followed
by a recovery phase at 25 ◦C. Though SA did not enhance Pn under
normal conditions, it attenuated heat-induced Pn and Rubisco activation
declines, accelerating post-stress recovery. It also preserved PSII func-
tion and sustained HSP21 accumulation, reinforcing PSII stability and
thermotolerance (Wang et al., 2010).
In Agrostis stolonifera (creeping bentgrass), exogenous application of
100 nM KAR1 signicantly enhanced drought resilience under
controlled drought conditions (soil water potential maintained at −0.6
MPa for 10 days) by increasing leaf RWC, chlorophyll retention,
photosystem II efciency, proline accumulation and higher membrane
stability. Antioxidant enzyme activities, including APX2, CAT1, POD2,
and Cu/Zn-SOD, were markedly increased, with associated upregulation
of KAR-responsive genes (AFL1, KAI2, and MAX2), TFs (ABF3, DREB2A,
bHLH148, and MYB13), and antioxidant-related genes. Meanwhile,
chlorophyll degradation genes (PPH and Chl-PRX) were downregulated,
delaying senescence under drought conditions (Tan et al., 2023).
Overexpression of the KARRIKINS INSENSITIVE2 (KAI2) gene in
A. thaliana signicantly enhanced drought tolerance by promoting
anthocyanin biosynthesis, improving stomatal regulation, maintaining
membrane integrity, and supporting cuticle formation. Under drought
conditions, KAI2 mutants showed increased water loss, cuticular
permeability, and membrane damage, coupled with reduced anthocy-
anin levels and lower sensitivity to ABA, a key regulator of drought
response. Transcriptomic analyses revealed KAI2 involvement in path-
ways that integrate ABA signalling to optimize plant adaptation. These
insights highlight the potential of targeting KAI2-dependent pathways to
develop crops with improved drought resilience (Li et al., 2017).
Additionally, 1 nM KAR1 supplementation in growth medium to
S. sebiferum seedlings determined improved biomass, longer taproots,
and a higher number of lateral roots, indicating better adaptation to
stress. Metabolomic analysis revealed elevated levels of organic acids
and amino acids, contributing to redox homeostasis. Interestingly, KAR1
modulated the expression of ABA signalling genes, such as those coding
for SNF1-RELATED PROTEIN KINASE2.3 and ABI5, without increasing
endogenous ABA levels, demonstrating its potential role in stress alle-
viation via ABA-independent pathways (Shah et al., 2020).
In Helianthus annuus, 20 days after spraying leaves with 1 mmol l
-1
GABA, plant height and fresh/dry biomass signicantly improved,
counteracting the sharp declines observed in these parameters under
drought and heat stress conditions. GABA alleviated the negative im-
pacts of the two stresses by improving gas exchange parameters such as
stomatal function and transpiration, maintaining chlorophyll levels, and
enhancing PSII efciency (Fv/Fm), and, therefore, photosynthetic per-
formance. GABA enhanced water relations by signicantly increasing
RWC (+25%) and reducing EL, MDA and H₂O₂ content under both stress
conditions, improving membrane stability and oxidative stress toler-
ance. On the molecular level, GABA-treated plants showed upregulation
of genes responsive to stress, including HSP70, DHNs and LEA proteins,
and aquaporins (AQP), which are crucial for osmotic adjustment, water
transport, and cellular protection. SOD, APX and POD activities
increased signicantly in GABA-treated plants, effectively scavenging
ROS and mitigating oxidative damage under stress. Furthermore, GABA
enhanced the accumulation of proline, total soluble sugars, and
α
-tocopherol (vitamin E), contributing to a 12% increase in seed yield
under drought and heat stress conditions. In chloroplasts,
α
-tocopherol
functions as a critical antioxidant, mitigating lipid peroxidation and
promoting the stability of cellular components during high-temperature
conditions. It also facilitates the accumulation of 3
′
-phosphoadenosine
5
′
-phosphate (PAP), which stabilizes microRNAs (miRNAs) involved in
gene regulation under thermal stress (Foyer and Hanke, 2022;
Munn´
e-Bosch, 2019; Rimbach et al., 2010). This highlights GABA’s ef-
cacy as a potent biostimulant for enhancing biochemical resilience and
productivity under combined drought and heat stresses (Abdel Razik
et al., 2021). Indeed, under drought and heat stress conditions, where
stomatal closure restricts CO₂ uptake, GABA may offer signicant
physiological benets by supporting plant metabolism. Carillo (2018)
proposed that GABA can reactivate the Calvin cycle by supplying CO₂
through the decarboxylation of glutamate, serving as an alternative CO₂
source when gas exchange is limited. This mechanism helps plants
dissipate excess energy, preventing photooxidative damage caused by
high light intensity. Also, GABA is a highly efcient osmolyte, protecting
the function and structure of membranes and proteins. Its potent anti-
oxidant activity, exceeding that of proline and GB, further enhances
stress tolerance by mitigating oxidative damage by eliminating ROS like
H
2
O
2
, singlet oxygen and superoxide radicals. The GABA shunt con-
tributes to stress recovery by supplying NADH and succinate to the
tricarboxylic acid (TCA) cycle, which supports cellular respiration and
energy production. In conjunction with calcium, GABA regulates the
expression of genes encoding 14–3–3 proteins, which are crucial for
nitrogen and carbon metabolism. Additionally, endogenous GABA syn-
thesis through the decarboxylation of glutamate, catalyzed by glutamate
decarboxylase (GAD), helps maintain cellular pH under salinity stress,
consuming protons and buffering cytosolic acidosis (Carillo, 2018).
Collectively, these properties enable GABA to alleviate drought and heat
stress, promoting osmotic balance, metabolic stability, and photosyn-
thetic recovery even under prolonged stress conditions (Carillo, 2018).
Seckin et al. (2009) investigated the impact of exogenous mannitol
application on wheat roots under osmotic and ionic stress. They found
that after applying 100 mmol l
-1
mannitol for 24 h, antioxidant enzyme
activities in the roots were signicantly enhanced, thereby improving
tolerance to drought and salinity stress. Application of 1 nmol l
-1
Thuricin 17 (Th17) through root drenching, a bioactive compound
produced by Bacillus thuringiensis NEB17, showed signicant potential in
enhancing plant resilience to drought in Brassica napus (canola). In fact,
the Th17 treatment increased root length, surface area, and volume by
over 20% under severe drought, while also boosting leaf area and
biomass. under combined heat and drought stress, 10 pmol l
-1
Th17 was
even more effective, reecting stress-specic responses. The application
of Th17 at both concentrations enhanced CAT, POD and SOD antioxi-
dant activities, improved water uptake, and strengthened root systems.
Root drenching, especially when paired with seed treatment, emerged as
a sustainable strategy for mitigating stress-related growth declines,
supporting eco-friendly agricultural resilience under drought and/or
heat stresses (Nazari and Smith, 2023). In Brassica napus, daily appli-
cation of 2 mL of 20 mmol l
-1
glutamate signicantly boosted proline
levels, improving RWC and drought resilience under limited irrigation
(La et al., 2020). Glutamate modulates proline metabolism by down-
regulating proline dehydrogenase (PDH) expression and upregulating
Δ¹-pyrroline-5-carboxylate synthetase (P5CS), leading to elevated pro-
line concentrations in the phloem and xylem, thereby enhancing
drought tolerance (La et al., 2020). Indeed, proline is the most wide-
spread osmolyte, serving as a key component in plant stress responses. It
is amphiphilic, neutral, and capable of detoxifying ROS. It can preserve
the integrity and functionality of proteins and membranes, stabilise the
redox potential of cells, and stimulate specic sequences within the
ProDH promoter region, particularly the Proline Responsive Element
(ACTCAT), thereby enhancing the expression of genes associated with
stress responses (Carillo, 2018). Unlike GB, which accumulates pri-
marily in young tissues, proline is synthesised rapidly in mature tissues
P. Carillo
Plant Stress 16 (2025) 100802
12
at the onset of stress and degraded after the stress subsides, supplying
energy, carbon, and nitrogen for recovery. This spatial and temporal
discrepancy underscores the unique adaptive strategies associated with
the accumulation of proline under stress (Carillo et al., 2008). In Satureja
hortensis, the foliar application of a PH, containing 43.7% total protein,
50% amino acids, and 20% total nitrogen, was carried out twice at the
onset of the shooting stage, and four weeks later, using a concentration
of 2 mL l
-1
of water. It enhanced proline content, pigment levels, and
yield under drought stress, supporting better growth and quality. The
application of PH signicantly improved dry matter yield, essential oil
content, and essential oil yield, with respective increases of 22%, 31%,
and 57%. The highest oil yields were achieved under conditions of
moderate (75% FC) and severe (50% FC) water decit and PH treat-
ments (Rezaei-Chiyaneh et al., 2023). In Zea mays (maize), hydroponic
applications of plant PHs containing 12% amino acids signicantly
enhanced root and shoot development and mitigated heat and drought
stress effects (Vaseva et al., 2022). Similarly, C. annuum treated with PHs
through fertigation (1.5 ml L⁻¹) demonstrated improved growth and
resilience by maintaining chlorophyll content, enhancing antioxidant
defences, and stabilizing biomass under stress conditions (Agliassa et al.,
2021). In S. lycopersicum, weekly foliar sprays of PHs derived from
Malvaceae (PH1) or Fabaceae (PH2) species (3 ml L⁻¹) increased levels of
dipeptides such as Arg-Leu and PyroGlu-Val, which played crucial roles
in activating defence mechanisms, improving WUE, and maintaining
growth under heat and drought stress (Leporino et al., 2024). Func-
tioning as metabolic regulators, these dipeptides inuence enzymatic
activities crucial for carbon ow and nitrogen metabolism during pe-
riods of stress. Arg-Leu may play a signicant role in nitrogen redistri-
bution and osmolyte production, fortifying the plant’s resilience to
abiotic challenges. Similarly, PyroGlu-Val is can be associated with
enhanced oxidative stress management through its ability to neutralize
ROS. Under stress conditions, these dipeptides support alternative car-
bon metabolism pathways, a vital mechanism when photosynthetic ef-
ciency is impaired, while also activating adaptive signalling processes
that sustain cellular stability and metabolic function. Additionally,
biostimulant treatments with PH1 and PH2 appear to induce metabolic
adjustments, evidenced by a reduction in the production of certain
stress-related detoxifying agents such as (R)-S-lactoylglutathione, which
is involved in methylglyoxal detoxication (Hoque et al., 2016). This
shift suggests that dipeptides, together with other antioxidant molecules
like phenols, may take over detoxication roles, enabling the plant to
allocate resources more effectively towards growth and antioxidant
defence systems (Leporino et al., 2024).
Foliar spraying of biostimulant preparations has been proven even
more effective when combined with seed priming across various crops.
Together, these two treatments signicantly boost physiological resil-
ience, nutrient biofortication, and yield quality, offering scalable
strategies for managing drought. In rainfed wheat, seed priming with a
nutrient solution containing urea (20 g L
_−1
), FeSO₄ (50 mg l
-1
), ZnSO₄
(50 mg l
-1
), and silicon (20 mg l
-1
) signicantly improved root length,
volume, and dry mass. Foliar applications at the anthesis stage with urea
(40 g l
-1
), silicon (112 mg l
-1
), FeSO₄ (6 g l
-1
), and ZnSO₄ (4 g l
-1
) further
enhanced physiological traits, including chlorophyll content and carot-
enoids, while reducing MDA and oxidative damage. Zn and Fe bio-
fortication increased grain yield by 29.17% and 19.51%, respectively,
and protein content from 11.14% to 12.46% (Moradi and Siosemardeh,
2023). In linseed, priming with β-aminobutyric acid (BABA) at doses of
50–100 µmol l
-1
and foliar spraying under severe drought (25% soil
moisture) enhanced antioxidant enzyme activity (e.g., CAT and POX)
and proline accumulation while reducing oxidative stress. Therefore,
BABA treatment maintained antioxidant defence and drought tolerance,
allowing an increase in root and shoot dry weights and improving yield
components such as capsule numbers, seeds per capsule, and seed
weight, ensuring productivity even under severe water stress (Yasir
et al., 2023).
5.3. Comparative effectiveness, application challenges, and risks
Biostimulants have been widely recognised for their role in
enhancing plant resilience to heat and drought stress, as extensively
discussed in the previous sections. However, while their efcacy is well-
established, a deeper understanding of their common mechanisms of
action, limitations, and potential risks is necessary for optimising their
application in real-world agricultural settings. This section synthesises
the comparative effectiveness of different biostimulants, highlights
shared physiological and molecular mechanisms and critically examines
their use’s challenges and potential risks.
5.3.1. Comparative effectiveness of biostimulants in mitigating heat and
drought stress
Different classes of biostimulants (e.g., algal extracts, microbial in-
oculants, PHs, and plant-derived metabolites) have been extensively
described for their role in mitigating heat and drought stress. Despite
differences in composition and mode of action, biostimulants share
several key physiological and molecular mechanisms that enhance plant
tolerance to heat and drought stress (Table 3). One of the most universal
mechanisms is regulating osmotic balance by accumulating compatible
solutes, such as proline, GB, and soluble sugars, which maintain cell
turgor and protect macromolecules under water decit conditions. Algal
extracts and protein hydrolysates (PHs) are particularly effective in this
regard, as they promote osmolyte synthesis, improve RWC, and stabilise
photosynthetic activity during stress. For instance, wheat seeds primed
with Spirulina platensis extract exhibited increased photosynthetic ef-
ciency and higher spike number under drought stress (Elnajar et al.,
2024). Similarly, PHs derived from legumes can improve stomatal
regulation and enhance leaf RWC in tomato and spinach (Leporino et al.,
2024). Another widely observed mechanism is enhancing antioxidant
defence systems to mitigate oxidative stress caused by heat and drought.
ROS accumulate rapidly under these conditions, leading to cellular
damage and impaired physiological functions. Microbial inoculants,
especially AMF and PGPR, have demonstrated the strongest effect in
enhancing ROS detoxication through increased activity of antioxidant
enzymes. Inoculation with Rhizophagus irregularis improved drought
tolerance in lettuce and tomato by enhancing shoot biomass (in lettuce
and tomato) and upregulating drought-responsive genes involved in
antioxidant defence (Ruiz-Lozano et al., 2016). Algal extracts, such as
Ulva lactuca, have also been reported to increase antioxidant enzyme
activities in Allium cepa, contributing to higher chlorophyll content and
root development under water stress (Muhie et al., 2020). Hormonal
regulation also plays a central role in biostimulant-mediated stress
tolerance. AMF and PGPR signicantly inuence ABA and SL biosyn-
thesis, which, in turn, regulate stomatal closure, root growth, and
drought-adaptive metabolic shifts. For example, in tomato plants treated
with R. irregularis, ABA biosynthesis genes (SlNCED2) and SL biosyn-
thesis genes (SlCCD7) were upregulated, promoting improved drought
resilience (Ruiz-Lozano et al., 2016). Conversely, PHs and plant extracts
primarily modulate JA and SA pathways, supporting cell membrane
stabilisation and osmolyte accumulation (Han et al., 2024). These
hormone-mediated responses contribute to better WUE, increased
thermotolerance, and overall stress adaptation. At the molecular level,
biostimulants inuence the expression of stress-responsive genes,
though their efcacy varies. PHs and algal extracts have been shown to
upregulate HSPs, DHNs, and LEA proteins, which protect cellular
structures from heat-induced damage (Gonz´
alez-Morales et al., 2021).
Treatments with Bacillus cereus in tomato plants induced the transcrip-
tion of heat-responsive TFs (SlHsfA1a) and autophagy-related genes
(SlATG5), leading to enhanced thermotolerance and root development
(Khan et al., 2020).
5.3.2.Application. challenges and practical limitations
While biostimulants demonstrate promising stress-mitigating effects
in controlled studies, several factors limit their effectiveness in eld
P. Carillo
Plant Stress 16 (2025) 100802
13
conditions. The Application method is critical in determining efcacy,
yet each technique has inherent constraints.
Seed priming is widely recognised for enhancing germination and
early stress tolerance by preconditioning seeds to activate protective
physiological mechanisms. However, its long-term effectiveness is
inconsistent, as bioactive compounds may degrade over time or be
insufciently retained in plant tissues, limiting sustained stress protec-
tion (Herrmann et al., 2024). The efcacy of priming is highly
species-dependent, requiring tailored protocols to optimise benets and
minimise potential adverse effects. For instance, variations in response
have been observed even among different cultivars of the same species,
like Brassica napus, emphasising the need for genotype-specic optimi-
zation of priming treatments (Zhu et al., 2021). Additionally, storage
conditions play a crucial role in maintaining priming-induced benets.
Studies highlight that primed seeds stored at 25 ◦C show reduced
viability and seedling vigour over time, underscoring the importance of
controlled storage environments to preserve priming efcacy (Hussain
et al., 2015). Another critical factor is the potential phytotoxicity of
some priming agents, particularly at high concentrations. For example,
the application of silver nanoparticles for seed priming has been re-
ported to inhibit germination and seedling growth in some plant species,
highlighting the need for careful concentration adjustments to prevent
toxicity (Budhani et al., 2019). Emerging technologies, such as
nano-priming, offer promising alternatives by improving the bioavail-
ability of priming agents while maintaining seed viability and stress
resilience. These approaches integrate nanotechnology-based carriers to
enhance the retention and controlled release of active compounds,
potentially overcoming the limitations of traditional priming methods
(Gohari et al., 2024). Despite these advancements, further research is
needed to rene priming techniques for large-scale agricultural appli-
cations, ensuring consistency across different environmental conditions
and crop species. Standardizing priming protocols, optimizing storage
conditions, and evaluating long-term eld performance remain crucial
steps for effectively integrating seed priming into climate-resilient
cropping systems.
As previously seen, foliar spraying is another widely adopted tech-
nique for delivering biostimulants directly to plant tissues, providing
rapid physiological benets by modulating stomatal regulation, anti-
oxidant defences, and hormonal signalling. However, its efcacy de-
pends on environmental conditions, formulation properties, and plant
surface characteristics, leading to variable outcomes in eld applica-
tions. Temperature, humidity, and wind speed signicantly inuence
foliar absorption. Under high temperatures and low humidity, rapid
evaporation reduces the retention time of sprayed solutions, limiting
their uptake by leaf tissues. At the same time, excessive rainfall may
wash off biostimulants before absorption occurs, resulting in inconsis-
tent effectiveness in open-eld conditions (Fern´
andez et al., 2021). The
physicochemical properties of foliar-applied biostimulants, including
molecular weight, polarity, and solubility, also play a key role in
determining absorption efciency. Low-molecular-weight compounds
like amino acids and phytohormones penetrate the cuticle more ef-
ciently than larger macromolecules like polysaccharides and proteins
(Sch¨
onherr, 2006). Additionally, leaf surface characteristics, including
cuticle thickness, trichome density, and wax composition, further in-
uence uptake rates. Younger leaves, with a thinner cuticle and higher
metabolic activity, generally absorb foliar-applied compounds more
effectively than mature leaves with a well-developed wax layer
(Fern´
andez et al., 2017). These factors contribute to variability in bio-
stimulant efcacy across different plant species and growth stages.
Advanced formulation strategies have been explored to address these
limitations to enhance foliar uptake and retention. Using nanocarriers,
such as lipid-based or polymeric nanoparticles, improves adhesion and
Table 3
Comparative effects of different classes of biostimulants in enhancing plant resilience to heat and drought stress.
Biostimulant Type Crop Species Mechanism of Action Physiological and Molecular Adaptations References
Algal Extracts (Ascophyllum
nodosum, Ulva lactuca, Spirulina
spp.)
Triticum aestivum (bread
wheat)
Enhanced osmotic adjustment via
increased GB and proline
accumulation
Improved photosynthetic efciency, increased
spike number and economic yield under
drought stress
Elnajar et al., 2024
Oryza sativa (rice) Increased antioxidant defense and
root elongation
Improved germination and root growth under
salinity and heat stress
Sornchai et al., 2014
Medicago truncatula
(alfalfa)
Modulation of auxin and cytokinin
homeostasis
Increased biomass, enhanced root-microbe
interactions, better drought resilience
Gitau et al., 2021
Microbial Biostimulants
(Rhizophagus irregularis, Bacillus
spp., Enterobacter spp.)
Lactuca sativa (lettuce),
Solanum lycopersicum
(tomato)
Arbuscular mycorrhizal (AM)
symbiosis, upregulation of ABA
biosynthesis genes
Increased AM colonization and improved
shoot dry weight (both in lettuce and tomato)
Ruiz-Lozano et al.,
2016
Triticum aestivum (bread
wheat)
Induction of ROS detoxication via
antioxidant enzyme activity
Increased proline levels, improved
antioxidant defenses, enhanced drought
tolerance
Shafque et al., 2023
Hordeum vulgare (barley) Enhanced root development,
nitrogen assimilation
Increased chlorophyll content, improved root
growth under drought
Suarez et al., 2015
Protein Hydrolysates (PHs) Zea mays (maize) Activation of osmoprotectant
pathways, improved nutrient
uptake
Increased root and shoot development,
improved drought resilience
Vaseva et al., 2022
Capsicum annuum
(pepper)
Stimulation of antioxidant enzyme
activity
Increased chlorophyll content, improved
biomass retention under drought stress
Agliassa et al., 2021
Solanum lycopersicum
(tomato)
Regulation of stomatal
conductance, upregulation of WUE-
related genes
Activation of defense-related dipeptides,
enhanced resilience to heat and drought
stress
Leporino et al., 2024
γ-Aminobutyric Acid (GABA) Helianthus annuus
(sunower)
Enhanced membrane stability,
reduced oxidative stress
Increased RWC, decreased electrolyte leakage,
MDA, and H₂O₂ levels, improved drought and
heat stress tolerance
Abdel Razik et al.,
2021
Silicon-Based Biostimulants Solanum lycopersicum
(tomato)
Improved water-use efciency,
stabilization of photosynthesis
Increased fruit yield and irrigation water
productivity under moderate drought stress
Chakma et al., 2021
Karrikins (KARs) Agrostis stolonifera
(creeping bentgrass)
Regulation of drought-induced
stress genes
Increased chlorophyll retention, improved
PSII efciency
Tan et al., 2023
Oryza sativa (rice) Enhancement of ABA-independent
pathways, improved stomatal
regulation
Delayed senescence, better water
conservation under heat stress
Pantoja-Benavides
et al., 2021
Polyamine-Based Biostimulants Triticum aestivum (bread
wheat)
Increased osmolyte accumulation,
regulation of stress-responsive
genes
Improved photosynthesis, reduced oxidative
damage, enhanced drought tolerance
Chaudhary et al., 2013
P. Carillo
Plant Stress 16 (2025) 100802
14
enables the controlled release of biostimulants, reducing losses due to
wash-off and evaporation (Kisvarga et al., 2022; Singh et al., 2024).
Surfactants and adjuvants are commonly incorporated to lower surface
tension, promote uniform droplet spreading, and increase contact time
on the leaf surface, thus enhancing penetration into the cuticle. How-
ever, the compatibility of these additives with different biostimulant
compounds must be carefully assessed to prevent potential phytotoxicity
or adverse chemical interactions (Silva et al., 2024). Moreover, despite
the advantages of foliar application, certain limitations must be
considered, particularly regarding dosage optimisation and systemic
translocation. Unlike soil-applied biostimulants, which can be absorbed
through roots and transported via the xylem, foliar-applied compounds
primarily exert localized effects, with limited movement to untreated
plant organs. Consequently, repeated applications throughout the
growth cycle are often required to sustain benecial effects, increasing
labour and operational costs, particularly in large-scale agricultural
systems (Niu et al., 2021; ´
Swierczy´
nski and Bosiacki, 2022). Future
research should optimise biostimulant formulations to enhance foliar
absorption while minimizing environmental losses. In addition, stan-
dardized protocols for foliar application, considering environmental
variability and plant-specic responses, will be essential for maximizing
the effectiveness of this biostimulant delivery approach.
Organic matter plays a crucial role in enhancing microbial activity,
improving soil structure, and increasing nutrient retention, thereby
promoting the sustained release and efcacy of biostimulants (du Jar-
din, 2015). Soil drenching is widely used to deliver microbial inoculants.
However, it is also an effective method for applying non-microbial
biostimulants, including algal extracts, protein hydrolysates, and
plant-derived metabolites, ensuring direct root uptake and systemic ef-
fects. However, its success depends on multiple soil properties, such as
pH, organic matter content, and moisture availability, inuencing bio-
stimulant stability and bioavailability (Rouphael and Colla, 2020). Mi-
crobial inoculants applied through soil drenching often face competition
from native soil microbiota, which can reduce colonisation success and
persistence. Additionally, abiotic factors such as temperature uctua-
tions, desiccation, and soil salinity can negatively impact microbial
viability and activity (O’Callaghan et al., 2022). For non-microbial
biostimulants, soil composition plays a critical role in determining ef-
cacy. Sandy soils with low water retention may lead to rapid leaching
of applied compounds, reducing their residence time in the rhizosphere
(Swartz and Motis, 2021). Conversely, clay-rich soils with high cation
exchange capacity can retain biostimulants for extended periods, but
this may limit their immediate bioavailability to plant roots (Fageria and
Moreira, 2011). Various formulation and application strategies have
been explored to overcome these challenges and improve biostimulant
efcacy in soil-drenching applications. For microbial inoculants, using
protective carriers, co-inoculation with compatible microbial consortia,
and optimising application timing can enhance survival and activity in
eld conditions (Aloo et al., 2022; Basu et al., 2021). For non-microbial
biostimulants, the incorporation of organic amendments such as bio-
char, compost, and humic substances has been shown to improve
retention, bioavailability, and nutrient-use efciency, supporting plant
growth under stress conditions (Canellas et al., 2015). These approaches
contribute to developing more sustainable and effective soil bio-
stimulant application strategies in both controlled and open-eld envi-
ronments.An additional signicant limitation to the widespread
adoption of biostimulants is the evaluation of their cost-effectiveness.
While these products offer agronomic benets, such as enhancing
plant growth and resilience to abiotic stresses, their economic justi-
cation is often more apparent in high-value crops like vegetables, fruits,
and vineyards, where higher prot margins can offset the investment.
For instance, a meta-analysis reported an average yield increase of
17.9% across various biostimulant categories, with the highest impact
observed in vegetable cultivation (e.g., greenhouse horticulture) (Li
et al., 2022). However, for extensive crops typical of Mediterranean
regions, such as cereals and legumes, the economic sustainability of
biostimulant use is less clear. The cost of these products and the actual
increase in productivity are central elements in adoption decisions.
Moreover, the variability of results in open-eld conditions, inuenced
by factors such as soil quality and climatic conditions, makes it chal-
lenging to quantify the direct benets derived from biostimulant use
precisely (Meena et al., 2025). Therefore, it is essential for farmers to
carefully assess the cost-benet ratio of biostimulants within their spe-
cic operational contexts. Further research and eld trials are necessary
to provide concrete data on the effectiveness and protability of these
products across diverse Mediterranean agricultural settings (Li et al.,
2022). In addition, exploring the interactive effects and potential syn-
ergies of biostimulants and biopesticides could offer a more integrated
approach to productivity and sustainability, particularly for organic
farming. A recent study currently underway at the University of Naples
(https://www.bbhort.unina.it/) is investigating these synergistic ef-
fects, aiming to enhance crop resilience, reduce chemical input de-
pendency, and promote long-term agricultural sustainability.
5.3.3.Biostimulants’. potential risks and regulatory considerations
Despite their largely benecial effects, the widespread adoption of
biostimulants raises important ecological and regulatory concerns,
particularly regarding microbial inoculants. One of the primary risks is
the potential for unintended environmental consequences, such as dis-
ruptions in native microbial communities and alterations in soil nutrient
dynamics. Introducing non-native microbial strains may outcompete
indigenous populations, affecting plant-microbe interactions and
potentially leading to imbalances in ecosystem function
(Barros-Rodríguez et al., 2020). Another key concern is contamination
with opportunistic pathogens, which could pose risks to plant health
and, in some cases, human safety. The large-scale application of mi-
crobial biostimulants necessitates rigorous quality control measures to
prevent contamination and ensure strain purity (Kumari et al., 2022).
The European Regulation (EU) 2019/1009 (EU, 2019) has established
safety guidelines for microbial biostimulants, but enforcement and
standardization remain challenging across different agricultural regions.
Additionally, concerns have been raised regarding horizontal gene
transfer between introduced and native microbial populations, which
could lead to unforeseen genetic modications with potential ecological
implications (Moore et al., 2022). Beyond microbial biostimulants,
biochemical-based products such as protein hydrolysates (PHs), algal
extracts, and plant-derived metabolites are generally considered safe,
with minimal direct toxicity to humans and the environment (Gupta and
van Staden, 2021). These biostimulants are primarily composed of
naturally occurring compounds, including amino acids, peptides, phy-
tohormones, and polysaccharides, which are biodegradable and typi-
cally do not persist in the environment at harmful levels (du Jardin,
2015). However, their safety prole depends on formulation properties,
concentration, and application methods. Biochemical-based bio-
stimulants, such as PHs and algal extracts, are generally considered safe
for plants, humans, and the environment. However, concerns regarding
potential contamination and unintended toxic effects have been raised.
One major concern is the overaccumulation of specic bioactive com-
pounds, such as auxins, cytokinins, or abscisic acid analogues, which at
high concentrations, may interfere with endogenous hormonal regula-
tion in plants. For example, excessive foliar application of auxin-rich
seaweed extracts has been associated with abnormal root elongation,
altered leaf morphology, and delayed owering in certain crop species
(Khan et al., 2009). Similarly, PHs rich in free amino acids can disrupt
nitrogen metabolism if applied at inappropriate doses, leading to
excessive vegetative growth at the expense of reproductive development
(Colla et al., 2017). Algal-based biostimulants may contain bioactive
secondary metabolites that could inuence non-target organisms,
including benecial insects and soil microfauna, though these effects
remain poorly characterised (Kapoore et al., 2021). Certain marine algae
can bioaccumulate heavy metals from their environment, which can
introduce contaminants into agricultural systems if not properly sourced
P. Carillo
Plant Stress 16 (2025) 100802
15
and processed. For instance, macroalgae such as Macrocystis pyrifera,
commonly used in commercial biostimulants, can accumulate metals
depending on the water quality of their habitat. Therefore, ensuring that
algal biomass is harvested from non-contaminated waters and un-
dergoes rigorous quality control is essential to prevent the introduction
of heavy metals into the soil and plant tissues (L¨
ahteenm¨
aki-Uutela
et al., 2021). In addition, the extraction process is critical in determining
the purity and safety of biostimulant formulations. Cold extraction
methods, such as Cold State Cell Disruption (CSCD), which do not
involve chemical solvents, can preserve the integrity of bioactive mol-
ecules while minimising the risk of contamination with heavy metals or
toxic by-products (Szparaga et al., 2023). Environmental conditions also
inuence the potential for phytotoxicity of biostimulants. Under high
temperatures or drought stress, foliar-applied biostimulants may cause
cuticular damage, leaf scorching, or osmotic imbalances, mainly if sur-
factants or adjuvants are included to enhance penetration (Silva et al.,
2024). Additionally, biostimulants that modify soil microbial activity,
such as polysaccharide-based formulations or humic substances, could
alter soil pH or disrupt benecial microbial communities if applied in
excess (Povero et al., 2016). Repeated applications of carbon-rich bio-
stimulants may further shift microbial populations, sometimes creating
favorable conditions for opportunistic or pathogenic microorganisms
(Pascale et al., 2019). To minimize potential risks, future research
should prioritise optimizing formulations, dening crop- and
environment-specic dosage guidelines, and strengthening regulatory
frameworks. Comprehensive toxicity assessments, including ecotoxico-
logical evaluations of soil and aquatic ecosystems, will ensure the safe
and sustainable integration of non-microbial biostimulants into agri-
cultural systems.
6. Conclusions
Biostimulants are increasingly recognized as a valuable strategy for
enhancing plant resilience to heat and drought stress, particularly in
Mediterranean cropping systems where climate variability poses sig-
nicant challenges. Their ability to improve water use efciency, miti-
gate oxidative damage, and support key physiological and biochemical
processes highlights their potential in sustainable agriculture. However,
their efcacy depends on multiple factors, including crop species,
environmental conditions, and application techniques, making their
optimization essential for broader adoption. Further research is needed
to rene evaluation methods, ensuring a better understanding of their
long-term effects on plant metabolism, soil microbial communities, and
overall ecosystem balance. Investigating the combined impact of bio-
stimulants under multiple stress conditions could reveal synergistic
benets while improving efciency and reducing input requirements.
Economic considerations remain central, requiring the development of
cost-effective solutions tailored to different agricultural systems. Addi-
tionally, strengthening regulatory frameworks will be key to ensuring
product reliability and environmental safety, fostering condence in
their practical use. By addressing these challenges, biostimulants can be
more effectively integrated into climate-resilient farming strategies,
contributing to sustainable food production while mitigating the
adverse effects of an increasingly unpredictable climate.
Funding
This work was supported by the projects BBHORT (PRIN 2022 PNRR
– grant P2022P52XK) and GREENHORT (PRIN 2022 - grant
2022WHTNZT).
CRediT authorship contribution statement
Petronia Carillo: Writing – review & editing, Writing – original
draft, Visualization, Validation, Supervision, Funding acquisition,
Conceptualization.
Declaration of competing interest
The author declares that she has no known competing nancial in-
terests or personal relationships that could have appeared to inuence
the work reported in this paper.
Data availability
No data was used for the research described in the article.
References
Abdel Razik, E.S., Alharbi, B.M., Pirzadah, T.B., Alnusairi, G.S.H., Soliman, M.H.,
Hakeem, K.R, 2021. γ-Aminobutyric acid (GABA) mitigates drought and heat stress
in sunower (Helianthus annuus L.) by regulating its physiological, biochemical and
molecular pathways. Physiol. Plant 172 (2), 505–527. https://doi.org/10.1111/
ppl.13216.
Agho, C., Avni, A., Bacu, A., Bakery, A., Balazadeh, S., Baloch, F.S., Bazakos, C.,
ˇ
Cerekovi´
c, N., Chaturvedi, P., Chauhan, H., De Smet, I., Dresselhaus, T., Ferreira, L.,
Fíla, J., Fortes, A.M., Fotopoulos, V., Francesca, S., García-Perez, P., Gong, W.,
Fragkostefanakis, S., 2024. Integrative approaches to enhance reproductive
resilience of crops for climate-proof agriculture. Plant Stress, 100704. https://doi.
org/10.1016/j.stress.2024.100704.
Agliassa, C., Mannino, G., Molino, D., Cavalletto, S., Contartese, V., Bertea, C.M.,
Secchi, F., 2021. A new protein hydrolysate-based biostimulant applied by
fertigation promotes relief from drought stress in Capsicum annuum L. Plant Physiol.
Biochem. 166, 1076–1086. https://doi.org/10.1016/j.plaphy.2021.07.015.
Agroberichten Buitenland, 2024. Spain: what happened in 2023 and what we can expect
in 2024. Dutch Ministry of Agriculture. Nature Food Qual. https://www.agroberich
tenbuitenland.nl/actueel/nieuws/2024/01/19/spain-what-happened-in-2023-and
-what-we-can-expect-in-2024 (accessed 13 February 2025). In.
Ahmed, M., Ullah, H., Attia, A., Tisarum, R., Cha-um, S., Datta, A., 2023. Interactive
effects of ascophyllum nodosum seaweed extract and silicon on growth, fruit yield
and quality, and water productivity of tomato under water stress. Silicon 15 (5),
2263–2278. https://doi.org/10.1007/s12633-022-02180-x.
Ahmed, N., Zhang, Y., Yu, H., Zhang, M., Zhou, Y., Li, Z., Gaber, A., 2019. Seed priming
with glycine betaine improve seed germination characteristics and antioxidant
capacity of wheat (triticum aestivum l.) seedlings under water-stress conditions.
Appl. Ecol. Environ. Res. 17. https://doi.org/10.15666/aeer/1704_83338350.
Allende-Montalban, R., Gabriel, J.L., de Andr´
es, E.F., Porcel, M.´
A., Santín-Montanya, M.
I., Gandía, M.L., Martín-Lammerding, D., Nieto, M.T., Delgado, M.d.M., San-Juan-
Heras, R., Tenorio, J.L., 2024. Nitrogen fertilization and sowing date as wheat
climate change adaptation tools under Mediterranean conditions. Eur. J. Agronomy
161, 127346. https://doi.org/10.1016/j.eja.2024.127346.
Aloo, B.N., Tripathi, V., Makumba, B.A., Mbega, E.R., 2022. Plant growth-promoting
rhizobacterial biofertilizers for crop production: the past, present, and future
[Review]. Front. Plant Sci. 13. https://doi.org/10.3389/fpls.2022.1002448.
Annunziata, M.G., Ciarmiello, L.F., Woodrow, P., Dell’Aversana, E., Carillo, P., 2019.
Spatial and Temporal Prole of Glycine Betaine Accumulation in Plants Under
Abiotic Stresses. Front. Plant Sci. 10, 230. https://doi.org/10.3389/fpls.2019.00230.
Annunziata, M.G., Ciarmiello, L.F., Woodrow, P., Maximova, E., Fuggi, A., Carillo, P.,
2017. Durum wheat roots adapt to salinity remodeling the cellular content of
nitrogen metabolites and sucrose [Article]. Front. Plant Sci. 7. https://doi.org/
10.3389/fpls.2016.02035. Article 2035.
Aragü´
es, R., Medina, E.T., Zribi, W., Clavería, I., ´
Alvaro-Fuentes, J., Faci, J., 2015. Soil
salinization as a threat to the sustainability of decit irrigation under present and
expected climate change scenarios. Irrig. Sci. 33 (1), 67–79. https://doi.org/
10.1007/s00271-014-0449-x.
Barghi, A., Jung, H.W., 2024. Insights into Bacillus zanthoxyli HS1-mediated systemic
tolerance: multifunctional implications for enhanced plant tolerance to abiotic
stresses. Physiol. Plant 176 (4), e14458. https://doi.org/10.1111/ppl.14458.
Barros-Rodríguez, A., Rangseekaew, P., Lasudee, K., Pathom-Aree, W., Manzanera, M.,
2020. Regulatory risks associated with bacteria as biostimulants and biofertilizers in
the frame of the European Regulation (EU) 2019/1009. Sci. Total. Environ. 740,
140239. https://doi.org/10.1016/j.scitotenv.2020.140239.
Basu, A., Prasad, P., Das, S.N., Kalam, S., Sayyed, R.Z., Reddy, M.S., El Enshasy, H., 2021.
Plant Growth Promoting Rhizobacteria (PGPR) as green bioinoculants: recent
developments, constraints, and prospects. Sustainability. 3, 13.
Begcy, K., Sandhu, J., Walia, H., 2018. Transient heat stress during early seed
development primes germination and seedling establishment in rice [Original
Research]. Front. Plant Sci. 9. https://doi.org/10.3389/fpls.2018.01768.
Bekele, Z., Asefa, A., 2023. Decit irrigation and mulching impacts on major crop yield
and water efciency: a review. Hydrology. https://doi.org/10.11648/j.
hyd.20231104.11.
Benitez-Alfonso, Y., Soanes, B.K., Zimba, S., Sinanaj, B., German, L., Sharma, V.,
Bohra, A., Kolesnikova, A., Dunn, J.A., Martin, A.C., Khashi u Rahman, M., Saati-
Santamaría, Z., García-Fraile, P., Ferreira, E.A., Fraz˜
ao, L.A., Cowling, W.A.,
Siddique, K.H.M., Pandey, M.K., Farooq, M., Foyer, C.H., 2023. Enhancing climate
change resilience in agricultural crops. Current Biol. 33 (23), R1246–R1261. https://
doi.org/10.1016/j.cub.2023.10.028.
Biju, S., Fuentes, S., Gupta, D., 2017. Silicon improves seed germination and alleviates
drought stress in lentil crops by regulating osmolytes, hydrolytic enzymes and
P. Carillo
Plant Stress 16 (2025) 100802
16
antioxidant defense system. Plant Physiol. Biochem. 119, 250–264. https://doi.org/
10.1016/j.plaphy.2017.09.001.
Blue Community, 2024. Italy: Water war in Sicily. Blue Community.
https://blue-community.net/2024/12/02/italy-water-war-in-sicily/ (accessed 8
December 2024).
Boini, A., Manfrini, L., Morandi, B., Corelli Grappadelli, L., Predieri, S., Daniele, G.M.,
L´
opez, G., 2021. High levels of shading as a sustainable application for mitigating
drought, in modern apple production. Agronomy (3), 11.
Brito, C., Dinis, L.-T., Moutinho-Pereira, J., Correia, C.M., 2019. Drought stress effects
and olive tree acclimation under a changing climate. Plants 8 (7).
Budhani, S., Egboluche, N.P., Arslan, Z., Yu, H., Deng, H., 2019. Phytotoxic effect of
silver nanoparticles on seed germination and growth of terrestrial plants. J. Environ.
Sci. Health C Environ. Carcinog. Ecotoxicol. Rev. 37 (4), 330–355. https://doi.org/
10.1080/10590501.2019.1676600.
Burato, A., Fusco, G.M., Pentangelo, A., Nicastro, R., Modugno, A.F., Scotto di
Covella, F., Ronga, D., Carillo, P., Campi, P., Parisi, M., 2024. Regulated decit
irrigation to boost processing tomato sustainability and fruit quality. Sustainability.
(9), 16.
Campobenedetto, C., Agliassa, C., Mannino, G., Vigliante, I., Contartese, V., Secchi, F.,
Bertea, C.M., 2021. A biostimulant based on seaweed (Ascophyllum nodosum and
Laminaria digitata) and yeast extracts mitigates water stress effects on tomato
(Solanum lycopersicum L.). Agriculture 11 (6), 557. https://www.mdpi.com
/2077-0472/11/6/557.
Campobenedetto, C., Grange, E., Mannino, G., van Arkel, J., Beekwilder, J., Karlova, R.,
Garabello, C., Contartese, V., Bertea, C.M., 2020. A biostimulant seed treatment
improved heat stress tolerance during cucumber seed germination by acting on the
antioxidant system and glyoxylate cycle. Front. Plant Sci. 11, 836. https://doi.org/
10.3389/fpls.2020.00836.
Canellas, L.P., Olivares, F.L., Aguiar, N.O., Jones, D.L., Nebbioso, A., Mazzei, P.,
Piccolo, A., 2015. Humic and fulvic acids as biostimulants in horticulture. Sci.
Hortic. 196, 15–27. https://doi.org/10.1016/j.scienta.2015.09.013.
Carbon Brief, 2024. State of the Climate: 2024 Will Be First Year Above 1.5◦C of Global
Warming. Carbon Brief. https://www.carbonbrief.
org/state-of-the-climate-2024-will-be-rst-year-above-1-5c-of-global-warming/ (accessed
8 December 2024).
Carillo, P., 2018. GABA shunt in durum wheat [Review]. Front. Plant Sci. 9. https://doi.
org/10.3389/fpls.2018.00100.
Carillo, P., Avice, J.-C., Vasconcelos, M.W., du Jardin, P., Brown, P.H., 2025.
Biostimulants in agriculture: editorial. Physiol. Plant 177 (1), e70046. https://doi.
org/10.1111/ppl.70046.
Carillo, P., Ciarmiello, L.F., Woodrow, P., Corrado, G., Chiaiese, P., Rouphael, Y., 2020a.
Enhancing sustainability by improving plant salt tolerance through macro- and
micro-algal biostimulants. Biology. (Basel) (9), 9.
Carillo, P., Ferrante, A., 2025. Decoding the intricate metabolic and biochemical changes
in plant senescence: a focus on chloroplasts and mitochondria. Ann. Bot. https://doi.
org/10.1093/aob/mcaf003.
Carillo, P., Kyratzis, A., Kyriacou, M.C., Dell’Aversana, E., Fusco, G.M., Corrado, G.,
Rouphael, Y., 2020b. Biostimulatory action of arbuscular mycorrhizal fungi
enhances productivity, functional and sensory quality in ‘Piennolo del Vesuvio’
cherry tomato landraces. Agronomy 10 (6).
Carillo, P., Mastrolonardo, G., Nacca, F., Parisi, D., Verlotta, A., Fuggi, A., 2008. Nitrogen
metabolism in durum wheat under salinity: accumulation of proline and glycine
betaine. Funct. Plant Biol. 35 (5), 412–426. https://doi.org/10.1071/fp08108.
Carlson, R., Tugizimana, F., Steenkamp, P.A., Dubery, I.A., Hassen, A.I., Labuschagne, N.,
2020. Rhizobacteria-induced systemic tolerance against drought stress in Sorghum
bicolor (L.) Moench. Microbiol. Res. 232, 126388. https://doi.org/10.1016/j.
micres.2019.126388.
Caruso, G., Formisano, L., Cozzolino, E., Pannico, A., El-Nakhel, C., Rouphael, Y.,
Tallarita, A., Cenvinzo, V., De Pascale, S., 2020. Shading affects yield, elemental
composition and antioxidants of perennial wall rocket crops grown from spring to
summer in Southern Italy. Plants 9 (8).
Cataldo, E., Eichmeier, A., Mattii, G.B., 2023. Effects of global warming on grapevine
berries phenolic compounds—a review. Agronomy (9), 13.
Chac´
on-Vozmediano, J.L., Martínez-Gascue˜
na, J., García-Romero, E., G´
omez-Alonso, S.,
García-Navarro, F.J., Jim´
enez-Ballesta, R., 2021. Effects of water stress on the
phenolic compounds of ‘merlot’ grapes in a semi-arid Mediterranean climate.
Horticulturae 7 (7).
Chakma, R., Saekong, P., Biswas, A., Ullah, H., Datta, A., 2021. Growth, fruit yield,
quality, and water productivity of grape tomato as affected by seed priming and soil
application of silicon under drought stress. Agric. Water. Manage 256, 107055.
https://doi.org/10.1016/j.agwat.2021.107055.
Chaudhary, D., Narula, N., Sindhu, S.S., Behl, R.K., 2013. Plant growth stimulation of
wheat (Triticum aestivum L.) by inoculation of salinity tolerant Azotobacter strains.
Physiol. Mol. Biol. Plants 19 (4), 515–519. https://doi.org/10.1007/s12298-013-
0178-2.
Chen, L., Xu, M., Zheng, Y., Men, Y., Sheng, J., Shen, L., 2014. Growth promotion and
induction of antioxidant system of tomato seedlings (Solanum lycopersicum L.) by
endophyte TPs-04 under low night temperature. Sci. Hortic. 176, 143–150. https://
doi.org/10.1016/j.scienta.2014.06.032.
Chen, S.Y., Zhang, X.Y., Pei, D., Sun, H.Y., Chen, S.L., 2007. Effects of straw mulching on
soil temperature, evaporation and yield of winter wheat: eld experiments on the
North China Plain. Ann. Appl. Biol. 150 (3), 261–268. https://doi.org/10.1111/
j.1744-7348.2007.00144.x.
Cocetta, G., Landoni, M., Pilu, R., Repiso, C., Nolasco, J., Alajarin, M., Ugena, L., Levy, C.
C.B., Scatolino, G., Villa, D., Ferrante, A., 2022. Priming treatments with
biostimulants to cope the short-term heat stress response: a transcriptomic prole
evaluation. Plants 11 (9), 1130. https://www.mdpi.com/2223-7747/11/9/1130.
Colla, G., Hoagland, L., Ruzzi, M., Cardarelli, M., Bonini, P., Canaguier, R., Rouphael, Y.,
2017. Biostimulant action of protein hydrolysates: unraveling their effects on plant
physiology and microbiome. Front. Plant Sci. 8, 2202. https://doi.org/10.3389/
fpls.2017.02202.
Copernicus, 2023. Floods in Northern Italy. Copernicus. May 2023. https://www.
copernicus.eu/en/media/image-day-gallery/oods-northern-italy-may-2023 (accessed 8
December 2024).
Copernicus, 2025. Summer 2023 was the hottest on record globally. Copernicus Climate
Change Service. https://climate.copernicus.eu/summer-2023-hottest-record (accessed
13 February 2025).
Cruz de Carvalho, M.H., 2008. Drought stress and reactive oxygen species: production,
scavenging and signaling. Plant Signal. Behav. 3 (3), 156–165. https://doi.org/
10.4161/psb.3.3.5536.
Cuin, T.A., Shabala, S., 2005. Exogenously supplied compatible solutes rapidly
ameliorate NaCl-induced potassium efux from barley roots. Plant Cell Physiol. 46
(12), 1924–1933. https://doi.org/10.1093/pcp/pci205.
De Santis, M.A., Giuzio, L., Tozzi, D., Soccio, M., Flagella, Z., 2024. Impact of no tillage
and low emission n fertilization on durum wheat sustainability, protability and
quality. Agronomy (12), 14.
De Vita, P., Di Paolo, E., Fecondo, G., Di Fonzo, N., Pisante, M., 2007. No-tillage and
conventional tillage effects on durum wheat yield, grain quality and soil moisture
content in southern Italy. Soil Tillage Res. 92 (1), 69–78. https://doi.org/10.1016/j.
still.2006.01.012.
Díaz-P´
erez, J.C., St. John, K., 2019. Bell Pepper (Capsicum annum L.) under colored
shade nets: plant growth and physiological responses. HortScience Horts 54 (10),
1795–1801. https://doi.org/10.21273/HORTSCI14233-19.
Dickinson, P.J., Kumar, M., Martinho, C., Yoo, S.J., Lan, H., Artavanis, G.,
Charoensawan, V., Sch¨
ottler, M.A., Bock, R., Jaeger, K.E., Wigge, P.A., 2018.
Chloroplast Signaling Gates Thermotolerance in <em>Arabidopsis</em>. Cell Rep.
22 (7), 1657–1665. https://doi.org/10.1016/j.celrep.2018.01.054.
dos Santos, T.B., Ribas, A.F., de Souza, S.G.H., Budzinski, I.G.F., Domingues, D.S., 2022.
Physiological responses to drought, salinity, and heat stress in plants: a review.
Stresses 2 (1), 113–135.
du Jardin, P., 2015. Plant biostimulants: denition, concept, main categories and
regulation. Sci. Hortic. 196, 3–14. https://doi.org/10.1016/j.scienta.2015.09.021.
Duan, B., Li, L., Chen, G., Su-Zhou, C., Li, Y., Merkeryan, H., Liu, W., Liu, X., 2021. 1-
aminocyclopropane-1-carboxylate deaminase-producing plant growth-promoting
rhizobacteria improve drought stress tolerance in grapevine (Vitis vinifera L.)
[Original Research]. Front. Plant Sci. 12. https://doi.org/10.3389/
fpls.2021.706990.
Duan, Y., Zhang, W., Li, B., Wang, Y., Li, K., Sodmergen, Han, Zhang, C., Y, Li, X., 2010.
An endoplasmic reticulum response pathway mediates programmed cell death of
root tip induced by water stress in Arabidopsis. New Phytologist 186 (3), 681–695.
https://doi.org/10.1111/j.1469-8137.2010.03207.x.
Duc, N.H., Csintalan, Z., Posta, K., 2018. Arbuscular mycorrhizal fungi mitigate negative
effects of combined drought and heat stress on tomato plants. Plant Physiol.
Biochem. 132, 297–307. https://doi.org/10.1016/j.plaphy.2018.09.011.
Dündar, G., Ramirez, V.E., Poppenberger, B., 2024. The heat shock response in plants:
new insights into modes of perception, signaling, and the contribution of hormones.
J. Exp. Bot. erae419. https://doi.org/10.1093/jxb/erae419.
EC, 2023. Inspirational ideas: Biostimulants for Sustainable Agriculture. European
Commission. https://ec.europa.eu/eip/agriculture/en/news/inspirational-ideas-bi
ostimulants-sustainable.html (accessed 12 December 2024).
EC, 2025. Crop Production in EU Countries – Cereals, rice, and Industrial crops.
European Commission. Eurostat. https://ec.europa.
eu/eurostat/databrowser/view/APRO_CPSH1__custom_7882816/bookmark/table?
lang=en&bookmarkId=8dd788d2-e08a-4412-b482-ed58dee4e072 (accessed 13
February 2025).
Edwards, R., Dixon, D.P., Walbot, V., 2000. Plant glutathione <em>S</em>-
transferases: enzymes with multiple functions in sickness and in health. Trends.
Plant Sci. 5 (5), 193–198. https://doi.org/10.1016/S1360-1385(00)01601-0.
Elnajar, M., Aldesuquy, H., Abdelmoteleb, M., Eltanahy, E., 2024. Mitigating drought
stress in wheat plants (Triticum Aestivum L.) through grain priming in aqueous
extract of spirulina platensis. BMC. Plant Biol. 24 (1), 233. https://doi.org/10.1186/
s12870-024-04905-z.
EU, 2019. Regulation (EU) 2019/1009 of the European Parliament and of the Council of
5 June 2019 laying down rules on the making available on the market of EU
fertilising products and amending Regulations (EC) No 1069/2009 and (EC) No
1107/2009 and repealing Regulation (EC) No 2003/2003. Available online.
https://eur-lex.europa.eu/eli/reg/2019/1009/oj (accessed on 26 December 2024). .
Euronews, 2023. Why do oods follow drought? Scientists show climate change is
fuelling more sudden shifts. Euronews. https://www.euronews.
com/green/2023/08/30/why-do-oods-fo
llow-drought-scientists-show-climate-change-is-fuelling-more-sudden-shifts (accessed 8
December 2024).
Fadiji, A.E., Babalola, O.O., Santoyo, G., Perazzolli, M., 2021. The potential role of
microbial biostimulants in the amelioration of climate change-associated abiotic
stresses on crops. Front. Microbiol. 12, 829099. https://doi.org/10.3389/
fmicb.2021.829099.
Fageria, N.K., & Moreira, A. (2011). Chapter four - the role of mineral nutrition on root
growth of crop plants. In D. L. Sparks (Ed.), Advances in Agronomy (Vol. 110, pp.
251–331). Academic Press. https://doi.org/10.1016/B978-0-12-385531-2.00004-9.
Fereres, E., Soriano, M.A., 2006. Decit irrigation for reducing agricultural water use.
J. Exp. Bot. 58 (2), 147–159. https://doi.org/10.1093/jxb/erl165.
P. Carillo
Plant Stress 16 (2025) 100802
17
Fernandes, ˆ
A., Chaski, C., Pereira, C., Kosti´
c, M., Rouphael, Y., Sokovi´
c, M., Barros, L.,
Petropoulos, S.A., 2022. Water stress alleviation effects of biostimulants on
greenhouse-grown tomato fruit. Horticulturae 8 (7).
Fern´
andez, V., Bahamonde, H.A., Javier Peguero-Pina, J., Gil-Pelegrín, E., Sancho-
Knapik, D., Gil, L., Goldbach, H.E., Eichert, T., 2017. Physico-chemical properties of
plant cuticles and their functional and ecological signicance. J. Exp. Bot. 68 (19),
5293–5306. https://doi.org/10.1093/jxb/erx302.
Fern´
andez, V., Gil-Pelegrín, E., Eichert, T., 2021. Foliar water and solute absorption: an
update. Plant J. 105 (4), 870–883. https://doi.org/10.1111/tpj.15090.
Formisano, L., 2022. Microclimate Factors Modulation Through the Use of Shading Nets
and Anti-insect Screens Impacts the Performance and Product Quality of Protected
Crops. University of Naples Federico II. http://www.fedoa.unina.
it/14543/1/Formisano_Luigi_34.pdf (accessed 13 February 2025).
Formisano, L., Ciriello, M., Cirillo, V., Pannico, A., El-Nakhel, C., Cristofano, F., Duri, L.
G., Giordano, M., Rouphael, Y., De Pascale, S., 2021. Divergent leaf morpho-
physiological and anatomical adaptations of four lettuce cultivars in response to
different greenhouse irradiance levels in early summer season. Plants 10 (6).
Foyer, C.H., Hanke, G., 2022. ROS production and signalling in chloroplasts:
cornerstones and evolving concepts. Plant J. 111 (3), 642–661. https://doi.org/
10.1111/tpj.15856.
Franzoni, G., Cocetta, G., Prinsi, B., Ferrante, A., Espen, L., 2022. Biostimulants on crops:
their impact under abiotic stress conditions. Horticulturae 8 (3).
Fusco, G.M., Nicastro, R., Rouphael, Y., Carillo, P., 2022. The effects of the microbial
biostimulants approved by EU Regulation 2019/1009 on yield and quality of
vegetable crops [review]. Foods. 11 (17). https://doi.org/10.3390/foods11172656.
Article 2656.
Galindo, A., Collado-Gonz´
alez, J., Gri˜
n´
an, I., Corell, M., Centeno, A., Martín-Palomo, M.
J., Gir´
on, I.F., Rodríguez, P., Cruz, Z.N., Memmi, H., Carbonell-Barrachina, A.A.,
Hern´
andez, F., Torrecillas, A., Moriana, A., P´
erez-L´
opez, D., 2018. Decit irrigation
and emerging fruit crops as a strategy to save water in Mediterranean semiarid
agrosystems. Agric. Water. Manage 202, 311–324. https://doi.org/10.1016/j.
agwat.2017.08.015.
Gallo, A.E., Christophe, A., Poupard, M., Boulord, R., Rolland, G., Prieto, J.A.,
Simonneau, T., Pallas, B., 2024. Effects of acclimation to long-term shading on
photosynthesis in grapevines: roles of non-structural carbohydrates and stomatal
conductance. Physiol. Plant 176 (6), e14636. https://doi.org/10.1111/ppl.14636.
Gedeon, S., Ioannou, A., Balestrini, R., Fotopoulos, V., Antoniou, C., 2022. Application of
biostimulants in tomato plants (Solanum lycopersicum) to enhance plant growth and
salt stress tolerance. Plants (22), 11.
Georgopoulou, E., Gakis, N., Voloudakis, D., Daskalakis, M., Saradis, Y., Lalas, D.P.,
Mirasgedis, S., 2024. Effectiveness of options for the adaptation of crop farming to
climate change in a country of the European South. Agriculture (10), 14.
Ghaleb, W., Barre, P., Teulat, B., Ahmed, L.Q., Escobar-Guti´
errez, A.J., 2022. Divergent
selection for seed ability to germinate at extreme temperatures in perennial ryegrass
(Lolium perenne L.) [Original Research]. Front. Plant Sci. 12. https://doi.org/
10.3389/fpls.2021.794488.
Gholami, R., Fahadi Hoveizeh, N., Zahedi, S.M., Gholami, H., Carillo, P., 2022. Effect of
three water-regimes on morpho-physiological, biochemical and yield responses of
local and foreign olive cultivars under eld conditions. BMC. Plant Biol. 22 (1), 477.
https://doi.org/10.1186/s12870-022-03855-8.
Gil, P.M., Lobos, P., Dur´
an, K., Olguín, J., Cea, D., Schaffer, B., 2018. Partial root-zone
drying irrigation, shading, or mulching effects on water savings, productivity and
quality of ‘Syrah’ grapevines. Sci. Hortic. 240, 478–483. https://doi.org/10.1016/j.
scienta.2018.06.050.
Gitau, M.M., Farkas, A., Balla, B., ¨
Ord¨
og, V., Fut´
o, Z., Mar´
oti, G., 2021. Strain-specic
biostimulant effects of chlorella and chlamydomonas green microalgae on Medicago
truncatula. Plants 10 (6), 1060. https://www.mdpi.com/2223-7747/10/6/1060.
Giuliani, M.M., Nardella, E., Gagliardi, A., Gatta, G., 2017. Decit irrigation and partial
root-zone drying techniques in processing tomato cultivated under Mediterranean
climate conditions. Sustainability. 9 (12).
Gohari, G., Jiang, M., Manganaris, G.A., Zhou, J., Fotopoulos, V., 2024. Next generation
chemical priming: with a little help from our nanocarrier friends. Trends. Plant Sci.
29 (2), 150–166. https://doi.org/10.1016/j.tplants.2023.11.024.
Gonz´
alez-Morales, S., Solís-Gaona, S., Vald´
es-Caballero, M.V., Ju´
arez-Maldonado, A.,
Loredo-Trevi˜
no, A., Benavides-Mendoza, A., 2021. Transcriptomics of biostimulation
of plants under abiotic stress. Front. Genet. 12, 583888. https://doi.org/10.3389/
fgene.2021.583888.
Gouache, D., Le Bris, X., Bogard, M., Deudon, O., Page, C., Gate, P., 2012. Evaluating
agronomic adaptation options to increasing heat stress under climate change during
wheat grain lling in France. Eur. J. Agronomy 39, 62. https://doi.org/10.1016/j.
eja.2012.01.009.
Grasso, M., Feola, G., 2012. Mediterranean agriculture under climate change: adaptive
capacity, adaptation, and ethics. Reg. Environ. Change 12 (3), 607–618. https://doi.
org/10.1007/s10113-011-0274-1.
Gupta, S., & van Staden, J. (2021). Biostimulants for crops from seed germination to plant
development- a practical approach.
Han, M., Kasim, S., Yang, Z., Deng, X., Saidi, N.B., Uddin, M.K., Shuib, E.M., 2024. Plant
extracts as biostimulant agents: a promising strategy for managing environmental
stress in sustainable agriculture. Phyton-Int. J. Exp. Botany 93 (9), 2149–2166.
https://doi.org/10.32604/phyton.2024.054009.
Heidari, M., Golpayegani, A., 2012. Effects of water stress and inoculation with plant
growth promoting rhizobacteria (PGPR) on antioxidant status and photosynthetic
pigments in basil (Ocimum basilicum L.). J. Saudi Soc. Agric. Sci. 11 (1), 57–61.
https://doi.org/10.1016/j.jssas.2011.09.001.
Herrmann, M.N., Grifn, L.G., John, R., Mosquera-Rodríguez, S.F., Nkebiwe, P.M.,
Chen, X., Yang, H., Müller, T., 2024. Limitations of soil-applied non-microbial and
microbial biostimulants in enhancing soil P turnover and recycled P fertilizer
utilization - a study with and without plants [Original Research]. Front. Plant Sci.
15. https://doi.org/10.3389/fpls.2024.1465537.
Hoque, T.S., Hossain, M.A., Mostofa, M.G., Burritt, D.J., Fujita, M., Tran, L.-S.P., 2016.
Methylglyoxal: an emerging signaling molecule in plant abiotic stress responses and
tolerance [Review]. Front. Plant Sci. 7. https://doi.org/10.3389/fpls.2016.01341.
Hossain, A., Teixeira Da Silva, J.A., 2012. Phenology, growth and yield of three wheat
(Triticum aestivum L.) varieties as affected by high temperature stress. Notulae
Scientia Biologicae 4 (3), 97–109. https://doi.org/10.15835/nsb437879.
Hussain, S., Zheng, M., Khan, F., Khaliq, A., Fahad, S., Peng, S., Huang, J., Cui, K., Nie, L.,
2015. Benets of rice seed priming are offset permanently by prolonged storage and
the storage conditions. Sci. Rep. 5, 8101. https://doi.org/10.1038/srep08101.
Iacuzzi, N., Tuttolomondo, T., Farruggia, D., Tortorici, N., Alaimo, F., De Santis, D.,
Rossini, F., Di Miceli, G., 2024. A two-year evaluation of biostimulant effects on yield
and quality parameters of tomato landrace ‘Pizzutello Delle Valli Ericine’ cultivated
without irrigation. J. Sustainable Agric. Environ. 3 (4). https://doi.org/10.1002/
sae2.70028.
Ibrahim, W.M., Ali, R.M., Hemida, K.A., Sayed, M.A., 2014. Role of Ulva lactuca extract
in alleviation of salinity stress on wheat seedlings. ScienticWorldJournal. 2014,
847290. https://doi.org/10.1155/2014/847290.
Icona Clima, 2023. Siccit`
a in Italia e crisi idrica in Sicilia: la politica deve assumersi la
responsabilit`
a di scelte urgenti. Icona Clima. https://www.iconaclima.it/italia/sic
cita-in-italia-e-crisi-idrica-in-sicilia-la-politica-deve-assumer
si-la-responsabilita-di-scelte-urgenti/ (accessed 8 December 2024).
IPCC. (2022). Climate Change 2022: impacts, adaptation and vulnerability, cross-chapter
Paper 4 (CCP4): Mediterranean Region. Intergovernmental Panel on Climate Change
(IPCC). https://www.ipcc.
ch/report/ar6/wg2/downloads/report/IPCC_AR6_WGII_CCP4.pdf (accessed 8 December
2024).
Jim´
enez-Arias, D., Hern´
andiz, A.E., Morales-Sierra, S., García-García, A.L., García-
Machado, F.J., Luis, J.C., Borges, A.A., 2022. Applying biostimulants to combat
water decit in crop plants: research and debate. Agronomy 12 (3).
Jim´
enez-Viveros, Y., Valiente-Banuet, J.I., 2023. Colored shading nets differentially
affect the phytochemical prole, antioxidant capacity, and fruit quality of piquin
peppers (Capsicum annuum L. var. glabriusculum). Horticulturae 9 (11).
Jovanovic, Z., Stikic, R., 2018. Partial root-zone drying technique: from water saving to
the improvement of a fruit quality [mini review]. Front. Sustain. Food Syst. 1.
https://doi.org/10.3389/fsufs.2017.00003.
Jumrani, K., Bhatia, V.S., 2014. Impact of elevated temperatures on growth and yield of
chickpea (Cicer arietinum L.). Field. Crops. Res. 164, 90–97. https://doi.org/
10.1016/j.fcr.2014.06.003.
Jagadeesha, Kakanur, Navathe, Y., Krishnappa, S., Ambati, G., Baviskar, D., Biradar, V.,
Magar, S., Mishra, N., Chandra, N., Mamrutha, Harohalli M., Govindan, V.,
Singh, Gyanendra P., Singh, G., 2024. Multi-environment analysis of nutritional and
grain quality traits in relation to grain yield under drought and terminal heat stress
in bread wheat and durum wheat. J. Agron. Crop. Sci. 210 (5), e12763. https://doi.
org/10.1111/jac.12763.
Kang, S., Zhang, J., 2004. Controlled alternate partial root-zone irrigation: its
physiological consequences and impact on water use efciency. J. Exp. Bot. 55
(407), 2437–2446. https://doi.org/10.1093/jxb/erh249.
Kapoore, R.V., Wood, E.E., Llewellyn, C.A., 2021. Algae biostimulants: a critical look at
microalgal biostimulants for sustainable agricultural practices. Biotechnol. Adv. 49,
107754. https://doi.org/10.1016/j.biotechadv.2021.107754.
Khan, M.A., Asaf, S., Khan, A.L., Jan, R., Kang, S.M., Kim, K.M., Lee, I.J., 2020. Extending
thermotolerance to tomato seedlings by inoculation with SA1 isolate of Bacillus
cereus and comparison with exogenous humic acid application. PLoS. One 15 (4),
e0232228. https://doi.org/10.1371/journal.pone.0232228.
Khan, W., Rayirath, U.P., Subramanian, S., Jithesh, M.N., Rayorath, P., Hodges, D.M.,
Critchley, A.T., Craigie, J.S., Norrie, J., Prithiviraj, B., 2009. Seaweed extracts as
biostimulants of plant growth and development. J. Plant Growth Regul. 28 (4),
386–399. https://doi.org/10.1007/s00344-009-9103-x.
KiliÇ, H., Tacettin, Y., 2010. The effect of drought stress on grain yield, yield components
and some quality traits of durum wheat (Triticum turgidum ssp. durum) Cultivars.
Not Bot Horti Agrobot Cluj Napoca 38. https://doi.org/10.15835/nbha3814274.
Kisvarga, S., Farkas, D., Boronkay, G., Nem´
enyi, A., Orl´
oci, L., 2022. Effects of
biostimulants in horticulture, with emphasis on ornamental plant production.
Agronomy 12 (5).
Koini, M.A., Alvey, L., Allen, T., Tilley, C.A., Harberd, N.P., Whitelam, G.C., Franklin, K.
A., 2009. High temperature-mediated adaptations in plant architecture require the
bHLH transcription factor PIF4. Current Biol. 19 (5), 408–413. https://doi.org/
10.1016/j.cub.2009.01.046.
Kostopoulou, E., Giannakopoulos, C., 2023. Projected changes in extreme wet and dry
conditions in Greece. Climate 11 (3), 49. https://www.mdpi.com/2225-115
4/11/3/49.
Kumari, M., Swarupa, P., Kesari, K.K., Kumar, A., 2022. Microbial inoculants as plant
biostimulants: a review on risk status. Life (Basel) 13 (1). https://doi.org/10.3390/
life13010012.
Kurniawan, A., Chuang, H.W., 2022. Rhizobacterial Bacillus mycoides functions in
stimulating the antioxidant defence system and multiple phytohormone signalling
pathways to regulate plant growth and stress tolerance. J. Appl. Microbiol. 132 (2),
1260–1274. https://doi.org/10.1111/jam.15252.
La, V.H., Lee, B.-R., Islam, M.T., Mamun, M.A., Park, S.-H., Bae, D.-W., Kim, T.-H., 2020.
Characterization of glutamate-mediated hormonal regulatory pathway of the
drought responses in relation to proline metabolism in Brassica napus L. Plants 9 (4).
Labuschagne, M.T., Elago, O., Koen, E., 2009. Inuence of extreme temperatures during
grain lling on protein fractions, and its relationship to some quality characteristics
P. Carillo
Plant Stress 16 (2025) 100802
18
in bread, biscuit, and durum wheat. Cereal. Chem. 86 (1), 61–66. https://doi.org/
10.1094/CCHEM-86-1-0061.
L¨
ahteenm¨
aki-Uutela, A., Rahikainen, M., Camarena-G´
omez, M.T., Piiparinen, J.,
Spilling, K., Yang, B., 2021. European Union legislation on macroalgae products.
Aquaculture Int. 29 (2), 487–509. https://doi.org/10.1007/s10499-020-00633-x.
Lazoglou, G., Papadopoulos-Zachos, A., Georgiades, P., Zittis, G., Velikou, K., Manios, E.
M., Anagnostopoulou, C., 2024. Identication of climate change hotspots in the
Mediterranean. Sci. Rep. 14 (1), 29817. https://doi.org/10.1038/s41598-024-
80139-1.
Lephatsi, M., Nephali, L., Meyer, V., Piater, L.A., Buthelezi, N., Dubery, I.A.,
Opperman, H., Brand, M., Huyser, J., Tugizimana, F., 2022. Molecular mechanisms
associated with microbial biostimulant-mediated growth enhancement, priming and
drought stress tolerance in maize plants. Sci. Rep. 12 (1), 10450. https://doi.org/
10.1038/s41598-022-14570-7.
Leporino, M., Rouphael, Y., Bonini, P., Colla, G., Cardarelli, M., 2024. Protein
hydrolysates enhance recovery from drought stress in tomato plants: phenomic and
metabolomic insights [Original Research]. Front. Plant Sci. 15. https://doi.org/
10.3389/fpls.2024.1357316.
Li, J., Van Gerrewey, T., Geelen, D., 2022. A meta-analysis of biostimulant yield
effectiveness in eld trials. Front. Plant Sci. 13, 836702. https://doi.org/10.3389/
fpls.2022.836702.
Li, N., Euring, D., Cha, J.Y., Lin, Z., Lu, M., Huang, L.-J., Kim, W.Y., 2021. Plant
hormone-mediated regulation of heat tolerance in response to global climate change
[review]. Front. Plant Sci. 11. https://doi.org/10.3389/fpls.2020.627969.
Li, W., Nguyen, K.H., Chu, H.D., Ha, C.V., Watanabe, Y., Osakabe, Y., Leyva-Gonz´
alez, M.
A., Sato, M., Toyooka, K., Voges, L., Tanaka, M., Mostofa, M.G., Seki, M., Seo, M.,
Yamaguchi, S., Nelson, D.C., Tian, C., Herrera-Estrella, L., Tran, L.P., 2017. The
karrikin receptor KAI2 promotes drought resistance in Arabidopsis thaliana. PLoS.
Genet. 13 (11), e1007076. https://doi.org/10.1371/journal.pgen.1007076.
Liu, J., Hasanuzzaman, M., Wen, H., Zhang, J., Peng, T., Sun, H., Zhao, Q., 2019. High
temperature and drought stress cause abscisic acid and reactive oxygen species
accumulation and suppress seed germination growth in rice. Protoplasma 256 (5),
1217–1227. https://doi.org/10.1007/s00709-019-01354-6.
Liu, X., Quan, W., Bartels, D., 2022. Stress memory responses and seed priming correlate
with drought tolerance in plants: an overview. Planta 255 (2), 45. https://doi.org/
10.1007/s00425-022-03828-z.
Lovisolo, C., Perrone, I., Carra, A., Ferrandino, A., Flexas, J., Medrano, H., Schubert, A.,
2010. Drought-induced changes in development and function of grapevine (Vitis
spp.) organs and in their hydraulic and non-hydraulic interactions at the whole-plant
level: a physiological and molecular update. Functional Plant Biol. 37 (2), 98–116.
https://doi.org/10.1071/FP09191.
Maxton, A., Singh, P., Masih, S.A., 2018. ACC deaminase-producing bacteria mediated
drought and salt tolerance in Capsicum annuum. J. Plant Nutr. 41 (5), 574–583.
https://doi.org/10.1080/01904167.2017.1392574.
Maytin, E.V., 1995. Heat shock proteins and molecular chaperones: implications for
adaptive responses in the skin. J. Invest. Dermatol. 104 (4), 448–455. https://doi.
org/10.1111/1523-1747.ep12605702.
Meena, D.C., Birthal, P.S., Kumara, T.M.K., 2025. Biostimulants for sustainable
development of agriculture: a bibliometric content analysis. Discover Agric. 3 (1), 2.
https://doi.org/10.1007/s44279-024-00149-5.
MFAT, 2023. Impact of drought on Spanish agriculture – October 2023. New Zealand
Ministry of Foreign Affairs and Trade. https://www.mfat.govt.nz/en/trade/mfat-ma
rket-reports/impact-of-drought-on-spanish-agriculture-october-2023 (accessed 13
February 2025). In.
Minoli, S., J¨
agermeyr, J., Asseng, S., Urfels, A., Müller, C., 2022. Global crop yields can
be lifted by timely adaptation of growing periods to climate change. Nat. Commun.
13 (1), 7079. https://doi.org/10.1038/s41467-022-34411-5.
Monde, Le, 2023. Why the Mediterranean basin has become the climate crisis ’hotspot’.
Le Monde. https://www.lemonde.fr/en/environment/article/2023/07/28/why-the-me
diterranean-basin-has-become-the-hotspot-of-climate-crisis_6070940_114.html (accessed
8 December 2024).
Moore, J.A.M., Abraham, P.E., Michener, Joshua K., Muchero, W., Cregger, Melissa A.,
2022. Ecosystem consequences of introducing plant growth promoting rhizobacteria
to managed systems and potential legacy effects. New Phytologist 234 (6),
1914–1918. https://doi.org/10.1111/nph.18010.
Moradi, L., Siosemardeh, A., 2023. Combination of seed priming and nutrient foliar
application improved physiological attributes, grain yield, and biofortication of
rainfed wheat [Original Research]. Front. Plant Sci. 14. https://doi.org/10.3389/
fpls.2023.1287677.
Moral, L., Rharrabti, Y., Villegas, D., Royo, C., 2003. Evaluation of grain yield and its
components in durum wheat under mediterranean conditions: an ontogenic
approach. Agron. J. 95, 266–274. https://doi.org/10.2134/agronj2003.2660.
Mughunth, R.J., Velmurugan, S., Mohanalakshmi, M., Vanitha, K., 2024. A review of
seaweed extract’s potential as a biostimulant to enhance growth and mitigate stress
in horticulture crops. Sci. Hortic. 334, 113312. https://doi.org/10.1016/j.
scienta.2024.113312.
Muhie, S.H., ¨
Ozdamar, C., G¨
okdas¸, Z., Njie, E.S., Memis¸, N., & Demir, I. (2020). Effect of
solid matrix priming with seaweed extract on germination and seedling performance
of onion seeds under abiotic stress conditions.
Mulugeta, S.M., Rad´
acsi, P., 2022. Inuence of drought stress on growth and essential oil
yield of ocimum species. Horticulturae 8 (2), 175. https://www.mdpi.com/2
311-7524/8/2/175.
Munn´
e-Bosch, S., 2019. Vitamin E function in stress sensing and signaling in plants. Dev.
Cell 48 (3), 290–292. https://doi.org/10.1016/j.devcel.2019.01.023.
Nazari, M., Smith, D.L., 2023. Mitigation of drought or a combination of heat and
drought stress effects on canola by Thuricin 17, a PGPR-produced compound
[Original Research]. Front. Sustain. Food Syst. 7. https://doi.org/10.3389/
fsufs.2023.1237206.
Nemhauser, J.L., Mockler, T.C., Chory, J., 2004. Interdependency of brassinosteroid and
auxin signaling in Arabidopsis. PLoS. Biol. 2 (9), E258. https://doi.org/10.1371/
journal.pbio.0020258.
Niu, J., Liu, C., Huang, M., Liu, K., Yan, D., 2021. Effects of Foliar Fertilization: a Review
of Current Status and Future Perspectives. J. Soil. Sci. Plant Nutr. 21 (1), 104–118.
https://doi.org/10.1007/s42729-020-00346-3.
NOAA, 2024. During a year of extremes, carbon dioxide levels surge faster than ever.
National Oceanic and Atmospheric Administration (NOAA). https://www.noaa.
gov/news-release/during-year-of-extremes-carbon-dioxide-levels-surge-faster-than-ever
(accessed 8 December 2024).
O’Callaghan, M., Ballard, R.A., Wright, D., 2022. Soil microbial inoculants for
sustainable agriculture: limitations and opportunities. Soil. Use Manage 38 (3),
1340–1369. https://doi.org/10.1111/sum.12811.
O’Donnell, E., Nogueira, L., Walters, C.G., Peterson, E.W.F., Irmak, S., 2023. Economics
of decit irrigation utilizing soil moisture probes in the western corn belt. Agric.
Resour. Econ. Rev. 52 (3), 563–581. https://doi.org/10.1017/age.2023.16.
Oancea, F., Velea, S., Fatu, V., Mincea, C., Ilie, L., 2013. Micro-algae based plant
biostimulant and its effect on water stressed tomato plants. Romanian J. Plant
Protect. 6, 104–117.
Oracz, K., Karpi´
nski, S., 2016. Phytohormones signaling pathways and ROS involvement
in seed germination [mini review]. Front. Plant Sci. 7. https://doi.org/10.3389/
fpls.2016.00864.
Osakabe, Y., Osakabe, K., Shinozaki, K., Tran, L.-S., 2014. Response of plants to water
stress [Mini Review]. Front. Plant Sci. 5. https://doi.org/10.3389/fpls.2014.00086.
Pandey, P., Ramegowda, V., Senthil-Kumar, M., 2015. Shared and unique responses of
plants to multiple individual stresses and stress combinations: physiological and
molecular mechanisms [Review]. Front. Plant Sci. 6. https://doi.org/10.3389/
fpls.2015.00723.
Pantoja-Benavides, A.D., Garces-Varon, G., Restrepo-Díaz, H., 2021. Foliar growth
regulator sprays induced tolerance to combined heat stress by enhancing
physiological and biochemical responses in rice [original research]. Front. Plant Sci.
12. https://doi.org/10.3389/fpls.2021.702892.
Papa, S., Fusco, G.M., Ciriello, M., Formisano, L., Woo, S.L., De Pascale, S., Rouphael, Y.,
Carillo, P., 2022. Microbial and non-microbial biostimulants as innovative tools to
increase macro and trace element mineral composition of tomato and spinach
[article]. Horticulturae 8 (12). https://doi.org/10.3390/horticulturae8121157.
Article 1157.
Paparella, S., Araújo, S.S., Rossi, G., Wijayasinghe, M., Carbonera, D., Balestrazzi, A.,
2015. Seed priming: state of the art and new perspectives. Plant Cell Rep. 34 (8),
1281–1293. https://doi.org/10.1007/s00299-015-1784-y.
Parmar, P., Kumar, R., Neha, Y., Srivatsan, V., 2023. Microalgae as next generation plant
growth additives: functions, applications, challenges and circular bioeconomy based
solutions. Front. Plant Sci. 14, 1073546. https://doi.org/10.3389/
fpls.2023.1073546.
Parri, S., Romi, M., Hoshika, Y., Giovannelli, A., Dias, M.C., Piritore, F.C., Cai, G.,
Cantini, C., 2023. Morpho-physiological responses of three Italian Olive Tree (Olea
europaea L.) cultivars to drought stress. Horticulturae 9 (7).
Pascale, A., Proietti, S., Pantelides, I.S., Stringlis, I.A., 2019. Modulation of the root
microbiome by plant molecules: the basis for targeted disease suppression and plant
growth promotion. Front. Plant Sci. 10, 1741. https://doi.org/10.3389/
fpls.2019.01741.
Pereira dos Anjos Neto, A., Oliveira, G., Costa Mello, S., da Silva, M., Gomes-Junior, F.,
Novembre, A.D., Azevedo, R., 2020. Seed priming with seaweed extract mitigate
heat stress in spinach: effect on germination, seedling growth and antioxidant
capacity. Bragantia 79. https://doi.org/10.1590/1678-4499.20200127.
P´
erez-P´
erez, J.G., Dodd, I.C., Botía, P., 2012. Partial rootzone drying increases water-use
efciency of lemon Fino 49 trees independently of root-to-shoot ABA signalling.
Functional Plant Biol. 39 (5), 366–378. https://doi.org/10.1071/FP11269.
P´
erez-P´
erez, J.G., Puertolas, J., Albacete, A., Dodd, I.C., 2020. Alternation of wet and dry
sides during partial rootzone drying irrigation enhances leaf ethylene evolution.
Environ. Exp. Bot. 176, 104095. https://doi.org/10.1016/j.envexpbot.2020.104095.
Philandras, C.M., Nastos, P.T., Kapsomenakis, J., Douvis, K.C., Tselioudis, G., Zerefos, C.
S., 2011. Long term precipitation trends and variability within the Mediterranean
region. Nat. Hazards Earth Syst. Sci. 11 (12), 3235–3250. https://doi.org/10.5194/
nhess-11-3235-2011.
Piri, R., Moradi, A., Balouchi, H., Salehi, A., 2019. Improvement of cumin (Cuminum
cyminum) seed performance under drought stress by seed coating and biopriming.
Sci. Hortic. 257, 108667. https://doi.org/10.1016/j.scienta.2019.108667.
Povero, G., Mejia, J.F., Di Tommaso, D., Piaggesi, A., Warrior, P., 2016. A systematic
approach to discover and characterize natural plant biostimulants [Perspective].
Front. Plant Sci. 7. https://doi.org/10.3389/fpls.2016.00435.
Prerostova, S., Dobrev, P.I., Gaudinova, A., Knirsch, V., K¨
orber, N., Pieruschka, R.,
Fiorani, F., Brzobohatý, B., ˇ
cerný, M., Spichal, L., Humplik, J., Vanek, T., Schurr, U.,
Vankova, R., 2018. Cytokinins: their impact on molecular and growth responses to
drought stress and recovery in arabidopsis [original research]. Front. Plant Sci. 9.
https://doi.org/10.3389/fpls.2018.00655.
Radwan, A.M., Ahmed, E.A., Donia, A.M., Mustafa, A.E., Balah, M.A., 2023. Priming of
Citrullus lanatus var. Colocynthoides seeds in seaweed extract improved seed
germination, plant growth and performance under salinity conditions. Sci. Rep. 13
(1), 11884. https://doi.org/10.1038/s41598-023-38711-8.
Rasul, F., Gupta, S., Olas, J.J., Gechev, T., Sujeeth, N., Mueller-Roeber, B., 2021. Priming
with a seaweed extract strongly improves drought tolerance in arabidopsis. Int. J.
Mol. Sci. 22 (3). https://doi.org/10.3390/ijms22031469.
P. Carillo
Plant Stress 16 (2025) 100802
19
Raven, J.A., 1985. Tansley review no. 2: regulation of ph and generation of osmolarity in
vascular plants: a cost-benet analysis in relation to efciency of use of energy,
nitrogen and water. New. Phytol. 101 (1), 25–77. https://doi.org/10.1111/j.1469-
8137.1985.tb02816.x.
Rezaei-Chiyaneh, E., Mahdavikia, H., Alipour, H., Dolatabadian, A., Battaglia, M.,
Maitra, S., Harrison, M., 2023. Biostimulants alleviate water decit stress and
enhance essential oil productivity: a case study with savory. Sci. Rep. https://doi.
org/10.1038/s41598-022-27338-w.
Rimbach, G., Moehring, J., Huebbe, P., Lodge, J.K., 2010. Gene-regulatory activity of
alpha-tocopherol. Molecules. 15 (3), 1746–1761. https://doi.org/10.3390/
molecules15031746.
Roslan, M.A.M., Zulkii, N.N., Sobri, Z.M., Zuan, A.T.K., Cheak, S.C., Abdul Rahman, N.
A., 2020. Seed biopriming with P- and K-solubilizing Enterobacter hormaechei sp.
improves the early vegetative growth and the P and K uptake of okra (Abelmoschus
esculentus) seedling. PLoS One. 15 (7), e0232860. https://doi.org/10.1371/journal.
pone.0232860.
Rouphael, Y., Carillo, P., Ciriello, M., Formisano, L., El-Nakhel, C., Ganugi, P., Fiorini, A.,
Miras Moreno, B., Zhang, L., Cardarelli, M., Lucini, L., Colla, G., 2023. Copper boosts
the biostimulant activity of a vegetal-derived protein hydrolysate in basil: morpho-
physiological and metabolomics insights. Front. Plant Sci. 14, 1235686. https://doi.
org/10.3389/fpls.2023.1235686.
Rouphael, Y., Colla, G., 2020. Toward a sustainable agriculture through plant
biostimulants: from experimental data to practical applications. Agronomy (10), 10.
Ruiz-Lozano, J.M., Aroca, R., Zamarre˜
no, ´
A.M., Molina, S., Andreo-Jim´
enez, B.,
Porcel, R., García-Mina, J.M., Ruyter-Spira, C., L´
opez-R´
aez, J.A., 2016. Arbuscular
mycorrhizal symbiosis induces strigolactone biosynthesis under drought and
improves drought tolerance in lettuce and tomato. Plant Cell Environ. 39 (2),
441–452. https://doi.org/10.1111/pce.12631.
Sabella, E., Aprile, A., Negro, C., Nicolì, F., Nutricati, E., Vergine, M., Luvisi, A., De
Bellis, L., 2020. Impacld. Agronomy 10 (6).
Sadok, W., Lopez, J.R., Smith, K.P., 2021. Transpiration increases under high-
temperature stress: potential mechanisms, trade-offs and prospects for crop
resilience in a warming world. Plant Cell Environ. 44 (7), 2102–2116. https://doi.
org/10.1111/pce.13970.
Saha, D., Choyal, P., Mishra, U.N., Dey, P., Bose, B., Md, P., Gupta, N.K., Mehta, B.K.,
Kumar, P., Pandey, S., Chauhan, J., Singhal, R.K., 2022. Drought stress responses
and inducing tolerance by seed priming approach in plants. Plant Stress 4, 100066.
https://doi.org/10.1016/j.stress.2022.100066.
Saidi, Y., Finka, A., Goloubinoff, P., 2011. Heat perception and signalling in plants: a
tortuous path to thermotolerance. New Phytologist 190 (3), 556–565. https://doi.
org/10.1111/j.1469-8137.2010.03571.x.
Sandalio, L.M., Espinosa, J., Shabala, S., Le´
on, J., Romero-Puertas, M.C., 2023. Reactive
oxygen species- and nitric oxide-dependent regulation of ion and metal homeostasis
in plants. J. Exp. Bot. 74 (19), 5970–5988. https://doi.org/10.1093/jxb/erad349.
Sato, H., Mizoi, J., Shinozaki, K., Yamaguchi-Shinozaki, K., 2024. Complex plant
responses to drought and heat stress under climate change. Plant J. 117 (6),
1873–1892. https://doi.org/10.1111/tpj.16612.
Savchenko, T., Tikhonov, K., 2021. Oxidative stress-induced alteration of plant central
metabolism. Life (Basel) 11 (4). https://doi.org/10.3390/life11040304.
Savoi, S., Herrera, J., Carlin, S., Lotti, C., Bucchetti, B., Peterlunger, E., Castellarin, S.D.,
Mattivi, F., 2020. From grape berries to wines: drought impacts on key secondary
metabolites. OENO One 54, 569–582. https://doi.org/10.20870/oeno-
one.2020.54.3.3093.
Savvides, A., Ali, S., Tester, M., Fotopoulos, V., 2016. Chemical priming of plants against
multiple abiotic stresses: mission possible? Trends. Plant Sci. 21 (4), 329–340.
https://doi.org/10.1016/j.tplants.2015.11.003.
Sch¨
onherr, J., 2006. Characterization of aqueous pores in plant cuticles and permeation
of ionic solutes. J. Exp. Bot. 57 (11), 2471–2491. https://doi.org/10.1093/jxb/
erj217.
Shafque, S., Imran, M., Kang, S.-M., Khan, M.A., Asaf, S., Kim, W.-C., Lee, I.-J., 2023.
Seed Bio-priming of wheat with a novel bacterial strain to modulate drought stress in
Daegu, South Korea. Front. Plant Sci. 14, 1118941. https://doi.org/10.3389/
fpls.2023.1118941.
Shah, F.A., Wei, X., Wang, Q., Liu, W., Wang, D., Yao, Y., Hu, H., Chen, X., Huang, S.,
Hou, J., Lu, R., Liu, C., Ni, J., Wu, L., 2020. Karrikin improves osmotic and salt stress
tolerance via the regulation of the redox homeostasis in the oil plant Sapium
sebiferum. Front. Plant Sci. 11, 216. https://doi.org/10.3389/fpls.2020.00216.
Shakir, M., Bano, A., Arshad, M., 2012. Rhizosphere bacteria containing ACC-deaminase
conferred drought tolerance in wheat grown under semi-arid climate. Soil Environ.
31, 108–112.
Shintu, P.V., 2015. Phosphate solubilising bacteria (Bacillus polymyxa) - an effective
approach to mitigate drought in tomato (Lycopersicon esculentum Mill.). Trop. Plant
Res. 2.
Silva, M.d.G.C., Medeiros, A.O., Converti, A., Almeida, F.C.G., Sarubbo, L.A., 2024.
Biosurfactants: promising biomolecules for agricultural applications. Sustainability.
(1), 16.
Singh, A., Sharma, A., Singh, O., Rajput, V.D., Movsesyan, H., Minkina, T., Alexiou, A.,
Papadakis, M., Singh, R.K., Singh, S., Sousa, J.R., El-Ramady, H.R., Zulqar, F.,
Kumar, R., Al-Ghamdi, A.A., Ghazaryan, K., 2024. In-depth exploration of
nanoparticles for enhanced nutrient use efciency and abiotic stresses management:
present insights and future horizons. Plant Stress 14, 100576. https://doi.org/
10.1016/j.stress.2024.100576.
Singh, R.P., Ma, Y., Shadan, A., 2022. Perspective of ACC-deaminase producing bacteria
in stress agriculture. J. Biotechnol. 352, 36–46. https://doi.org/10.1016/j.
jbiotec.2022.05.002.
Sita, K., Sehgal, A., Kumar, J., Kumar, S., Singh, S., Siddique, K.H.M., Nayyar, H., 2017.
Identication of high-temperature tolerant lentil (Lens culinaris Medik.) genotypes
through leaf and pollen traits [Original Research]. Front. Plant Sci. 8. https://doi.
org/10.3389/fpls.2017.00744.
Sofo, A., Dichio, B., Xiloyannis, C., Masia, A., 2004. Lipoxygenase activity and proline
accumulation in leaves and roots of olive trees in response to drought stress. Physiol.
Plant 121 (1), 58–65. https://doi.org/10.1111/j.0031-9317.2004.00294.x.
Sornchai, P., Saithong, N., Srichompoo, Y., Unartngam, A., Iamtham, S., 2014. Effect of
Spirulina maxima aqueous extract on seed germination and seedling growth of mung
bean, Vigna radiata and rice, Oryza sativa var. Japonica. J. Int. Soc. Southeast Asian
Agric. Sci. 20, 77–84.
Spada, M., Marín-Sanz, M., Bigini, V., Quagliata, G., Coppa, E., Barro, F., Savatin, D.,
Ruggeri, R., Sestili, F., Rossini, F., Astol, S., 2024. Use of biostimulants for water
stress mitigation in two durum wheat (Triticum durum Desf.) genotypes with
different drought tolerance. Plant Stress 14, 100566. https://doi.org/10.1016/j.
stress.2024.100566.
Statuto, D., Picuno, P., Abdel-Ghany, A., 2019. Shading methods for crop protection
under greenhouse in Mediterranean areas.
Suarez, C., Cardinale, M., Ratering, S., Steffens, D., Jung, S., Montoya, A.M.Z., Geissler-
Plaum, R., Schnell, S., 2015. Plant growth-promoting effects of Hartmannibacter
diazotrophicus on summer barley (Hordeum vulgare L.) under salt stress. Appl. Soil
Ecol. 95, 23–30. https://doi.org/10.1016/j.apsoil.2015.04.017.
Sun, Y., Chen, S., Grin, I.R., Zharkov, D.O., Yu, B., Li, H., 2024. The dual role of
methylglyoxal in plant stress response and regulation of DJ-1 protein. Physiol. Plant
176 (6), e14608. https://doi.org/10.1111/ppl.14608.
Swartz, S., Motis, T., 2021. Crop responses to natural farming soil drenches on sandy
soils. From: ECHO Research Notes 2 (7). https://www.echocommunity.
org/en/resources/77df2c02-72d7-4e34-807a-34eee4242c26 (accessed 16 February
2025).
´
Swierczy´
nski, S., Bosiacki, M., 2022. The effect of foliar spray treatments with various
biostimulants and fertilisers on the growth of M.9 rootstock stoolings. Agronomy 12
(3).
Syed, A.R., 2020. Blue Gold: an emerging source of global conicts. modern diplomacy.
https://moderndiplomacy.
eu/2020/08/10/blue-gold-an-emerging-source-of-global-conicts/ (accessed 8 December
2024).
Szparaga, A., Kocira, S., Kapusta, I., Zaguła, G., 2023. Solid–liquid extraction of bioactive
compounds as a green alternative for developing novel biostimulant from Linum
usitatissimum L. Chem. Biol. Technol. Agric. 10 (1), 108. https://doi.org/10.1186/
s40538-023-00482-9.
Takahashi, F., Kuromori, T., Urano, K., Yamaguchi-Shinozaki, K., Shinozaki, K., 2020.
Drought stress responses and resistance in plants: from cellular responses to long-
distance intercellular communication [review]. Front. Plant Sci. 11. https://doi.org/
10.3389/fpls.2020.556972.
Tamindˇ
zi´
c, G., Ignjatov, M., Miljakovi´
c, D., ˇ
Cervenski, J., Miloˇ
sevi´
c, D., Nikoli´
c, Z.,
Vasiljevi´
c, S., 2023. Seed priming treatments to improve heat stress tolerance of
garden pea (Pisum sativum L.). Agriculture 13 (2).
Tan, Z.-Z., Wang, Y.-T., Zhang, X.-X., Jiang, H.-Y., Li, Y., Zhuang, L.-L., Yu, J.-J., Yang, Z.-
M., 2023. Karrikin1 enhances drought tolerance in creeping bentgrass in association
with antioxidative protection and regulation of stress-responsive gene expression.
Agronomy 13 (3).
The Independent, 2023. Sicily’s water crisis: how climate change is worsening drought in
the region. The Independent. https://www.independent.co.
uk/news/world/europe/sicily-water-climate-change-drought-b2602545.html (accessed 8
December 2024).
Thomas, A., Yadav, B.K., ˇ
Simůnek, J., 2024. Water uptake by plants under nonuniform
soil moisture conditions: a comprehensive numerical and experimental analysis.
Agric. Water. Manage 292, 108668. https://doi.org/10.1016/j.agwat.2024.108668.
Tiwari, S., Prasad, V., Chauhan, P.S., Lata, C., 2017. Bacillus amyloliquefaciens confers
tolerance to various abiotic stresses and modulates plant response to phytohormones
through osmoprotection and gene expression regulation in rice [original research].
Front. Plant Sci. 8. https://doi.org/10.3389/fpls.2017.01510.
Toki´
c, M., Leljak Levani´
c, D., Ludwig-Müller, J., Bauer, N., 2023. Growth and molecular
responses of tomato to prolonged and short-term heat exposure. Int. J. Mol. Sci. 24
(5). https://doi.org/10.3390/ijms24054456.
Tong, S.M., Xi, H.X., Ai, K.J., Hou, H.S., 2017. Overexpression of wheat TaNCED gene in
Arabidopsis enhances tolerance to drought stress and delays seed germination. Biol.
Plant. 61 (1), 64–72. https://doi.org/10.1007/s10535-016-0692-5.
Turan, M., Ekinci, M., Argin, S., Brinza, M., Yildirim, E., 2023. Drought stress
amelioration in tomato (Solanum lycopersicum L.) seedlings by biostimulant as
regenerative agent [Original Research]. Front. Plant Sci. 14. https://doi.org/
10.3389/fpls.2023.1211210.
UNOCHA. (2023). Horn of Africa drought: regional humanitarian overview & call to action
(Revised, 26 May 2023). United Nations Ofce for the Coordination of Humanitarian
Affairs (UNOCHA), 2023. https://www.unocha.
org/publications/report/ethiopia/horn-afric
a-drought-regional-humanitarian-overview-call-action-revised-26-may-2023 (accessed 12
December 2024).
Vaseva, I.I., Simova-Stoilova, L., Kostadinova, A., Yuperlieva-Mateeva, B.,
Karakicheva, T., Vassileva, V., 2022. Heat-stress-mitigating effects of a protein-
hydrolysate-based biostimulant are linked to changes in Protease, DHN, and HSP
gene expression in maize. Agronomy 12 (5), 1127. https://doi.org/10.3390/
agronomy12051127.
Vu, L.D., Gevaert, K., De Smet, I., 2019a. Feeling the heat: searching for plant
thermosensors. Trends. Plant Sci. 24 (3), 210–219. https://doi.org/10.1016/j.
tplants.2018.11.004.
P. Carillo
Plant Stress 16 (2025) 100802
20
Vu, L.D., Xu, X., Gevaert, K., De Smet, I., 2019b. Developmental plasticity at high
temperature1[OPEN]. Plant Physiol. 181 (2), 399–411. https://doi.org/10.1104/
pp.19.00652.
Wahbi, S., Wakrim, R., Aganchich, B., Tahi, H., Serraj, R., 2005. Effects of partial
rootzone drying (PRD) on adult olive tree (Olea europaea) in eld conditions under
arid climate: I. Physiological and agronomic responses. Agric. Ecosyst. Environ. 106
(2), 289–301. https://doi.org/10.1016/j.agee.2004.10.015.
Wahid, A., Shabbir, A., 2005. Induction of Heat stress tolerance in barley seedlings by
pre-sowing seed treatment with glycinebetaine. Plant Growth Regul. 46 (2),
133–141. https://doi.org/10.1007/s10725-005-8379-5.
Wang, L.-J., Fan, L., Loescher, W., Duan, W., Liu, G.-J., Cheng, J.-S., Luo, H.-B., Li, S.-H.,
2010. Salicylic acid alleviates decreases in photosynthesis under heat stress and
accelerates recovery in grapevine leaves. BMC. Plant Biol. 10 (1), 34. https://doi.
org/10.1186/1471-2229-10-34.
Wang, W., Vinocur, B., Shoseyov, O., Altman, A., 2004. Role of plant heat-shock proteins
and molecular chaperones in the abiotic stress response. Trends. Plant Sci. 9 (5),
244–252. https://doi.org/10.1016/j.tplants.2004.03.006.
Wang, W., Zhang, C., Zheng, W., Lv, H., Li, J., Liang, B., Zhou, W., 2022. Seed priming
with protein hydrolysate promotes seed germination via reserve mobilization,
osmolyte accumulation and antioxidant systems under PEG-induced drought stress.
Plant Cell Rep. 41 (11), 2173–2186. https://doi.org/10.1007/s00299-022-02914-6.
Wang, X., Zhang, J., Song, J., Huang, M., Cai, J., Zhou, Q., Dai, T., Jiang, D., 2020.
Abscisic acid and hydrogen peroxide are involved in drought priming-induced
drought tolerance in wheat (Triticum aestivum L.). Plant Biol. 22 (6), 1113–1122.
https://doi.org/10.1111/plb.13143.
WEF, 2022. The horn of Africa is facing an unprecedented drought. what is the world
doing to help solve it? World Economic Forum (WEF). 21 July 2022, updated 3 March
2023. https://www.weforum.
org/stories/2022/07/africa-drought-food-insecurity-starvation-somalia/ (accessed 8
December 2024).
WMO, 2023. Climate change indicators reached record levels in 2023. World Meteorol.
Organiz. https://wmo.
int/news/media-centre/climate-change-indicators-reached-record-levels-2023-wmo
(accessed 13 February 2025).
WMO, 2025. WMO conrms 2024 as warmest year on record at about 1.55◦C above pre-
industrial level. https://wmo.int/news/media-centre/wmo-conrms-2024-warm
est-year-record-about-155degc-above-pre-industrial-level (accessed 13 February 2025).
Xiang, Y., Huang, Y., Xiong, L., 2007. Characterization of stress-responsive CIPK genes in
rice for stress tolerance improvement. Plant Physiol. 144 (3), 1416–1428. https://
doi.org/10.1104/pp.107.101295.
Xu, Z., Sun, M., Jiang, X., Sun, H., Dang, X., Cong, H., Qiao, F., 2018. Glycinebetaine
biosynthesis in response to osmotic stress depends on jasmonate signaling in
watermelon suspension cells [original research]. Front. Plant Sci. 9. https://doi.org/
10.3389/fpls.2018.01469.
Yasir, T.A., Ateeq, M., Wasaya, A., Hussain, M., Sarwar, N., Mubeen, K., Aziz, M.,
Iqbal, M.A., Ogbaga, C.C., Al-Ashkar, I., Rahman, M.A., Sabagh, A.E., 2023. Seed
priming and foliar supplementation with β-aminobutyric acid alleviates drought
stress through mitigation of oxidative stress and enhancement of antioxidant defense
in linseed (Linum usitatissimum L.). Phyton-Int. J. Exper. Botany 92 (11),
3113–3131. https://doi.org/10.32604/phyton.2023.029502.
Zarrouk, O., Brunetti, C., Egipto, R., Pinheiro, C., Genebra, T., Gori, A., Lopes, C.M.,
Tattini, M., Chaves, M.M., 2016. Grape ripening is regulated by decit irrigation/
elevated temperatures according to cluster position in the canopy [original
research]. Front. Plant Sci. 7. https://doi.org/10.3389/fpls.2016.01640.
Zhang, Z., Qiang, H., McHugh, A., Li, H., Wang, Q., Lu, Z., 2016. Effect of conservation
farming practices on soil organic matter and stratication in a mono-cropping
system of Northern China. Soil Tillage Res. 156, 173–181. https://doi.org/10.1016/
j.still.2015.10.008.
Zhou, M., Hassan, M.J., Peng, Y., Liu, L., Liu, W., Zhang, Y., Li, Z., 2021. γ-Aminobutyric
Acid (GABA) priming improves seed germination and seedling stress tolerance
associated with enhanced antioxidant metabolism, DREB expression, and dehydrin
accumulation in white clover under water stress. Front. Plant Sci. 12, 776939.
https://doi.org/10.3389/fpls.2021.776939.
Zhu, Z.H., Sami, A., Xu, Q.Q., Wu, L.L., Zheng, W.Y., Chen, Z.P., Jin, X.Z., Zhang, H.,
Li, Y., Yu, Y., Zhou, K.J., 2021. Effects of seed priming treatments on the germination
and development of two rapeseed (Brassica napus L.) varieties under the co-
inuence of low temperature and drought. PLoS. One 16 (9), e0257236. https://doi.
org/10.1371/journal.pone.0257236.
P. Carillo
Plant Stress 16 (2025) 100802
21