Skip to main content

Morpho-physiological mechanisms of two different quinoa ecotypes to resist salt stress

Abstract

Background

Quinoa (Chenopodium quinoa Willd.) is a facultative halophyte showing various mechanisms of salt resistance among different ecotype cultivars. This study aimed to determine salt resistance limits for a Peruvian sea level ecotype “Hualhuas” and a Bolivian salar ecotype “Real” and elucidate individual mechanisms conferring differences in salt resistance between these cultivars. The plants were grown in sandy soil and irrigated with various saline solutions concentrations (0, 100, 200, 300, 400, and 500 mM NaCl) under controlled conditions.

Results

High salinity treatment (500 mM NaCl) reduced the plant growth by 80% and 87% in Hualhuas and Real cultivars, respectively. EC50 (water salinity which reduces the maximum yield by 50%) was at a salinity of 300 mM NaCl for Hualhuas and between 100 and 200 mM NaCl for Real plants. Both cultivars were able to lower the osmotic potential of all organs due to substantial Na+ accumulation. However, Hualhuas plants exhibited distinctly lower Na+ contents and consequently a higher K+/Na+ ratio compared to Real plants, suggesting a more efficient control mechanism for Na+ loading and better K+ retention in Hualhuas plants. Net CO2 assimilation rates (Anet) were reduced, being only 22.4% and 36.2% of the control values in Hualhuas and Real, respectively, at the highest salt concentration. At this salinity level, Hualhuas plants showed lower stomatal conductance (gs) and transpiration rates (E), but higher photosynthetic water use efficiency (PWUE), indicative of an efficient control mechanism over the whole gas-exchange machinery.

Conclusion

These results reveal that Hualhuas is a promising candidate in terms of salt resistance and biomass production compared to Real.

Peer Review reports

Background

Water shortage and soil salinity are gaining great attention worldwide, due to their negative impacts on plant growth, crop yield, and thus food security, particularly in arid climates [1]. Nearly 8% of the earth’s surface and more than 30% of the global irrigated lands are salt-affected at significant levels [2]. The widespread soil salinization is becoming more prevalent, especially with the extension of intensive agriculture (to meet the future needs of humans) and the inappropriate use of limited water resources [3]. The problem is expected to get even worse over the next decades on the eve of global climatic changes [4, 5]. In this context, the implementation of adaptive measures to sustain crop productivity in salt-affected and marginal areas is a key priority [6]. Sustainable utilization of halophytes in salt-affected and/or degraded lands using saline water would be a feasible solution that fosters crop productivity in such areas [7, 8]. Among a number of underutilized halophytic species, Chenopodium quinoa (Family: Amaranthaceae), can be an excellent cash crop with tremendous potentials for marginal and salt-affected areas [9,10,11]. According to Bazile et al. [12], quinoa domestication is thought to have begun in the Andean region 7000 years ago. It is cultivated in various agro-ecological zones from 5° North Latitude in southern Colombia to 43° South Latitude in the Tenth Region of Chile and the Argentinean Andes, with altitudinal distribution ranges from sea level to 4000 m above sea level [13]. The large geographical distribution of quinoa accompanied by a great genetic diversity led to the identification of five quinoa ecotypes namely: salares (salt flats), highlands, inter-Andean valleys, yungas, and coastal lowlands [14]. Due to the broad diversity of its native habitats, quinoa is characterized by a marked variability of environmental adaptation, specifically to soil salinity [15,16,17,18,19,20], drought [21,22,23], frost [24], high solar radiation [25] and temperature [26]. Quinoa grains are rich in a wide range of important minerals (Ca, P, Mg, Fe, and Zn), vitamins (B1, B9, C, and E), oil (containing large amounts of linoleate, linolenate, and natural antioxidants), and protein-containing ample amounts of essential amino acids such as lysine and methionine [27,28,29,30,31]. Its potential as a nutritious and resistant crop was recognized by the Food and Agriculture Organization of the United Nations (FAO), which declared the year 2013 as the International Year of Quinoa [32]. Because of these characteristics, quinoa is thrust into the limelight as a non-conventional cash crop, especially in regions where salinity has been recognized as a major agricultural problem [33]. At present, quinoa is cultivated in more than 50 countries outside its origin, with some reports demonstrating an acceptable adaptation in the United States, Canada, Italy, Morocco, India, Pakistan, and Egypt [12]. Incorporation of this promising species into the Egyptian agricultural production system under non-permissive conditions (salinity) calls, however, for precise knowledge about its performance under salt stress, the limits of salinity resistance, and individual mechanisms enabling the plant to grow in saline habitats [16].

In general, the response of quinoa to salinity is characteristic of facultative halophytes, with plant growth stimulation occurring at low and moderate salinity levels [16, 20, 34]. Previous studies demonstrated the ability of some quinoa genotypes to survive even at seawater salinity (up to 50 dS m− 1) [10, 34]. Salt resistance of halophytic species is, in most cases, multi-genic, governed by an array of interconnected physiological, morphological, and biochemical mechanisms operating at cellular, organ, and whole plant levels [35]. These mechanisms are closely related to the four major constraints of salinity on plant growth, i.e., osmotic effects, nutritional imbalance, ion toxicity, and restriction of CO2 gas exchange [36, 37]. Quinoa has been reported to exhibit a wide range of salt resistance mechanisms. Regarding adaptation to osmotic stress, quinoa showed a very efficient system to adjust osmotically and to reduce its transpiration to maintain a positive water balance in response to salinity [34, 38]. Salinity resistance in quinoa has been attributed to a delicate balance between osmotic adjustment and ion (Na+, K+, and Cl) accumulation [15, 16, 20, 34]. Nonetheless, increases in organic osmolytes such as proline, glycine betaine, and soluble sugars have been also reported in quinoa [24, 39, 40].

A thorough review of the literature has shown that quinoa displays a high degree of genetic distancing. Its response to salinity stress is strongly genotype dependent, as revealed by many comparative studies on many different accessions, landraces, and cultivars [10, 15, 38, 41, 42]. As a response to salinity, plants have to control their transpiration through sensitive stomatal closure to avoid water loss [37]. Consequently, apparent assimilation rates decline due to restricted CO2 availability for carboxylation reactions (stomatal limitation of photosynthesis) [43]. Leaf gas exchange, stomatal conductance, and transpiration rates have been shown to decrease in quinoa under salinity [16, 20, 23, 44, 45]. The plant’s ability to maintain high CO2 assimilation rates at minimum H2O loss and energy consumption is crucial for its growth under saline conditions [46]. In many salt-resistant species, including quinoa, stomatal limitation of photosynthesis reduces the transpiration rate, leading to higher photosynthetic water use efficiency [20, 40, 47, 48]. While stomatal limitations of photosynthesis are considered the main cause of reduced photosynthetic rate under mild salinity stress, non-stomatal limitations (metabolic and diffusive impairments) become predominant as salinity stress intensifies [49].

Against this background, the present study was designed to screen and compare the eco-physiological responses of a Peruvian (sea level ecotype) quinoa cultivar “Hualhuas” and a Bolivian (salar ecotype) cultivar “Real” to water salinity under greenhouse conditions. These cultivars originate from different agroecological zones and are expected to exhibit various levels of adaptability to salt stress. Our intent was to determine the limits of salt resistance and the individual adaptive mechanisms conferring resistance differences in these cultivars. Comparing the responses of these closely related cultivars to saline irrigation may give an opportunity for elucidating the key mechanism(s) involved in salinity resistance in quinoa and open prospects for selecting the most suitable cultivar for comprehensive and commercial field trials under Egyptian conditions.

Results

Effect of salinity on plant growth and development

Comparative growth responses of Hualhuas and Real plants to varying salinity concentrations are illustrated in Fig. (1A and B). Phenotypic variations between different salinity treatments were visible four weeks after the beginning of salt treatments. Under control conditions, plants of Real cultivar showed relatively higher fresh weight (118.5 g/plant) compared to those of Hualhuas (112.4 g/plant) (Fig. 1A and B). Increasing NaCl salinity led to a progressive growth inhibition in both cultivars under evaluation, owing to gradual and significant (P < 0.05) reductions in the fresh weight of all plant organs (Fig. 1A and B). High salinity treatments (500 mM NaCl) drastically reduced the plant fresh weight by about 80% and 87% in Hualhuas and Real plants, respectively, relative to the corresponding controls (Fig. 1A and B). In both cultivars, salt-induced growth reduction was much more pronounced for the shoot compared to the root parts, leading to decline the shoot: root fresh weight ratio from 13 to 9 and from 13 to 1 for Hualhuas and Real plants, respectively. The salt resistance threshold (water salinity that causes initial significant reduction in the maximum expected yield) [50] was at salinity level of 100 mM NaCl for Hualhuas plants and at 200 mM NaCl for Real plants. EC50 was at a salinity of 300 mM NaCl for Hualhuas plants and between 100 and 200 mM NaCl for Real plants (Fig. 1A and B).

Fig. 1
figure 1

Plant development and growth responses of different organs (expressed as fresh weight) of C. quinoa cv. Hualhuas (A) and C. quinoa cv. Real (B) at various NaCl salinities. The dotted lines mark the EC50 values, while arrows show salinity resistance threshold. RFW, root fresh weight; SFW, stem fresh weight; LaFW, adult leaves fresh weight; LjFW, juvenile leaves fresh weight; InFW, inflorescence fresh weight. Each column represents the mean values of six replicates and the bars represent standard errors. Columns with the same letter are not significantly different at P < 0.05, according to Duncan’s multiple range test

Effect of salinity on water relations

Water content

Under control conditions, water content of Hualhuas plants was comparatively low, ranging between 77.0% (R) and 86.2% (La). On average over different plant organs, water content of Real plants was in the range of 78.58% (R) and 89.8% (La) (Fig. 2A and B). Transient increases in the water contents of all plant organs were observed as NaCl concentration in the external nutrient solution increased (Fig. 2A and B). Maximum water contents, ranging from 85 to 90% and 83–92% for Hualhuas and Real, respectively, were reached at water salinities between 200 and 400 mM NaCl (Fig. 2A and B). In Hualhuas plants, further increase in water salinity slightly reduced the root water content, but increased that of adult and juvenile leaves (Fig. 2A). As for Real plants, high salinity treatment slightly reduced the water content of all plant organs relative to their controls (Fig. 2B).

Fig. 2
figure 2

Effect of various NaCl salinity levels on water contents (WC in % FW) of different plant organs of Hualhuas (A) and Real (B) plants. R, root; La, adult leaves; Lj, juvenile leaves. Each column represents the mean values of six replicates and the bars represent standard errors. Columns with the same letter are not significantly different at P < 0.05, determined by Duncan’s multiple range test

Osmotic potential (ψs)

On average over different plant organs, ψs measured from − 0.25 MPa (R) to -1.24 MPa (Lj) and from − 0.42 MPa (R) to -1.03 MPa (Lj) in Hualhuas and Real plants, respectively, under control conditions (Fig. 3A and B). ψs fell gradually in all plant organs and became more negative with increasing water salinity (Fig. 3A and B). It reached from − 0.57 MPa (R) to -2.9 MPa (Lj) and from − 1.59 MPa (R) to -3.69 MPa (La) in Hualhuas and Real plants, respectively, when the plants were exposed to full strength water salinity (Fig. 3A and B).

Fig. 3
figure 3

Effect of increasing water salinity on osmotic potential of different plant organs of Hualhuas (A) and Real (B) plants. R, root; La, adult leaves; Lj, juvenile leaves. Each column represents the mean values of six replicates and the bars represent standard errors. Columns with the same letter are not significantly different at P < 0.05, determined by Duncan’s multiple range test

Effect of salinity on Na+, K+, and K+/Na+

Whatever the salinity treatment, Na+ concentrations in roots were lower than those of shoots in both cultivars. Elevating water salinity progressively and significantly increased Na+ concentrations, but decreased that of K+ in all plant organs of both cultivars under evaluation. This led to a gradual reduction in K+/ Na+ ratio in both cultivars (Fig. 4A and B). Salt-induced reduction in K+/Na+ ratio was higher in Real plants, as high salinity treatment significantly (P < 0.05) declined this ratio by roughly 13, 22, and 25 fold in the roots, adult leaves, and juvenile ones, respectively, compared to the corresponding controls (Fig. 4B).

Fig. 4
figure 4

Effect of increasing NaCl salinity on K+/Na+ ratio of different plant organs of Hualhuas (A) and Real (B) plants. R, root; La, adult leaves; Lj, juvenile leaves. Each column represents the mean values of six replicates and the bars represent standard errors. Columns with the same letter are not significantly different at P < 0.05, determined by Duncan’s multiple range test

Effect of salinity on proline concentration

Under control conditions, low proline concentrations were found in all plant organs, particularly, in the roots of both quinoa cultivars (Fig. 5A and B). On average, proline concentrations were between 2.89 µg g− 1 (R) and 29.55 µg g− 1 (Lj) in Hualhuas plants and between 2.16 µg g− 1 (R) and 19.65 µg g− 1 (La) in Real plants (Fig. 5A and B). In Hualhuas plants, proline concentrations enhanced markedly as NaCl salinity rose, with maximal increases of about 40%, 53%, and 40% in roots, adult leaves, and juvenile leaves at seawater salinity (Fig. 5A). The same trend was observed for Real plants, but only in the juvenile leaves (75% increases relative to the controls). On the contrary, proline concentrations in the roots and adult leaves were reduced by 12% and 47%, respectively, in this cultivar (Fig. 5B).

Fig. 5
figure 5

Effect of different water salinity levels on proline concentration (µg g− 1 FW) of different plant organs of Hualhuas (A) and Real (B) plants. R, root; La, adult leaves; Lj, juvenile leaves. Each column represents the mean values of six replicates and the bars represent standard errors. Columns with the same letter are not significantly different at P < 0.05, determined by Duncan’s multiple range test

Effect of salinity on photosynthetic pigment and H2O/CO2-gas exchange parameters

Chlorophyll (a), (b), and carotenoid concentrations

Chlorophyll (a) was the prominent pigment in the leaves of both cultivars, with 42.5 and 38.4 µg cm− 2 for Hualhuas and Real plants, respectively, under control conditions (Fig. 6A and B). It was gradually and significantly (P < 0.05) decreased as the external salinity increased. High water salinity led to reductions of 51.4% and 28.6% in Hualhuas and Real plants, respectively, compared to the respective controls (Fig. 6A and B). The same trend was also observed for Chl (b), as increasing water salinity lowered Chl (b) concentrations in both cultivars, although with a less severe effect compared to Chl (a) (Fig. 6A and B). Full-strength salinity resulted in significant reductions of about 46.7% and 22.4% in Hualhuas and Real, respectively, compared to controls. As a consequence, the ratio Chl (a)/Chl (b) distinctly declined from 3.3 to 3.0 (Hualhuas) and from 3.8 to 3.5 (Real) at the highest salinity treatment. Carotenoids concentration was significantly (P < 0.05) reduced in the leaves of Hualhuas plants, but slightly (statistically not significant) declined in those of Real plants in response to water salinity. High water salinity led to reduce their concentrations by about 47.3% and 29.8% in Hualhuas and Real, respectively (Fig. 6A and B).

Fig. 6
figure 6

Effect of different water salinity levels on the concentration of photosynthetic pigments of Hualhuas (A) and Real (B) plants. Each column represents the mean values of six replicates and the bars represent standard errors. Columns with the same letter are not significantly different at P < 0.05, determined by Duncan’s multiple range test

H2O/CO2-gas exchange parameters

The response of CO2 assimilation rates (Anet) to varying photosynthetic active radiation (PAR) is illustrated in Fig. (7). Regardless of salinity treatment, Anet was gradually increased with increasing PAR, then started to level off, reached a saturation plateau at PAR intensities of 872.3 and 1011.9 µmol m− 2s− 1 in Hualhuas and Real plants, respectively, under control conditions (Fig. 7; Table 1). Anet at light saturation measured about 13.5 and 15.1µmol CO2 m− 2s− 1 for Hualhuas and Real plants, respectively (Table 2). Anet of both cultivars was significantly (P < 0.05) and steadily reduced as the water salinity rose, reached only about 22.4 and 36.2 of the control values in Hualhuas and Real, respectively, at the highest salinity treatment (Table 2). This was accompanied by a decline of about 3.2 and 27.7% in the apparent carboxylation efficiency (ΦCO2) for Hualhuas and Real, respectively, at the highest water salinity level (Table 1). At this salinity level, the photosynthesis of both cultivars was also saturated at distinctly lower PAR intensities compared to respective controls (Fig. 7; Table 1). Additionally, the saturation irradiance (Ls) was decreased to 334.5 and 429.0 µmol m− 2s− 1 in Hualhuas and Real plants, respectively (Table 1). The light compensation point (Lc) was reduced as external salinity rose, being 27.4 and 45.2 µmol m− 2 s− 1 for Hualhuas and Real plants, respectively, at full strength salinity (Table 1). Dark respiration (Rd) also dropped by 45.3 and 33.4% in Hualhuas and Real plants at high salinity treatment (Tables 1 and 2). Salt-induced reduction in Anet was accompanied with a gradual and significant (P < 0.05) decrease in stomatal conductance (gs). High water salinity treatment led to reduce gs by about 93.5 and 77.9% in Hualhuas and Real, respectively (Table 2). This consequently inhibited the transpiration rate (E) by 95.3 and 80.2% in Hualhuas and Real, respectively, at salinity level of 500 mM NaCl (Table 2). In Hualhuas plants, the photosynthetic water use efficiency (PWUE) was steadily enhanced as the salinity rose, with more than three folds increments at 500 mM NaCl (Table 2). The same trend of salt-induced gradual enhancement in PWUE was observed for Real plants, but only up to 400 mM NaCl, where PWUE was increased by about two folds (Table 2). Higher salinity, however, led to a drastic reduction in PWUE by 88.8% in the plants of this cultivar (Table 2). Raising water salinity led to gradual reductions in Ci, which consequently lowered the ratio between internal and external CO2 concentrations (Ci/Ca) in Hualhuas plants (Table 2). The same tendency was also noted for Real plants, but again up to a salinity level of 400 mM NaCl; thereafter, Ci/Ca ratio was progressively increased to reach the same levels observed under control conditions (Table 2).

Table 1 Leaf gas exchange parameters calculated from A-PAR response curves of Hualhuas and Real plants as affected by water salinity. ΦCO2, apparent carboxylation efficiency (µmol CO2 µmol–1 photons); Lc, light compensation point (µmol photons m–2 s–1); Ls, light saturation point (µmol photons m–2 s–1); Rd, dark respiration rate (µmol CO2 m–2 s–1)
Table 2 Effect of different NaCl salinity treatments on the net assimilation rate (Anet), transpiration rate (E), photosynthetic water use efficiency (PWUE), stomatal conductance (gs) and the ratio between internal and external CO2 concentration (Ci/Ca) of Hualhuas and Real plants
Fig. 7
figure 7

Effect of NaCl salinity on light response curves of Hualhuas and Real plants. Anet, net photosynthetic rate (µmol m− 2 s− 1); PAR, photosynthetic active radiation (µmol m− 2 s− 1)

Discussion

The intention of this study was to determine the range of salt resistance and the individual mechanisms conferring differences in resistance between the Peruvian quinoa cultivar “Hualhuas” and the Bolivian cultivar “Real”. The growth and biomass accumulation of these cultivars were found to greatly differ depending on salinity level (Fig. 1A and B). Significant differences in plant growth rate, morphological and agronomical traits were previously reported in quinoa and have been attributed to its wide genetic variability [51].

Raising water salinity led to a substantial growth reduction, the response that obviously differed between the quinoa cultivars under evaluation. While low water salinity did not significantly alter the biomass of Real plants, it markedly reduced the fresh weight of Hualhuas plants. Similarly, Sanchez et al. [44] observed that total plant biomass of Real variety was not negatively influenced by water salinity up to 20 ds m− 1. Furthermore, several earlier studies reported an optimal growth in C. quinoa between 100 and 200 mM NaCl [16, 20]. In accordance with our results, Geissler et al. [47] found that low and moderate NaCl water salinities led to reduce the biomass of Hualhuas plants in a pot experiment carried out in an open-top chamber. High water salinity treatment, however, drastically reduced the plant biomass of both cultivars and again, the screened cultivars displayed considerable salt resistance variability. Plants of Hualhuas were least affected by high water salinity, exhibiting a growth reduction of about 80% relative to the controls, while plants of Real were more sensitive, with biomass reduction of about 87% relative to the controls. Similarly, the growth of the Peruvian cultivar Hualhuas [16, 47] and the Danish variety Titicaca [34] was significantly inhibited in response to high water salinity (50 ds m− 1). Here, the salt resistance threshold was at salinity levels of 10 and 20 ds m− 1 for Hualhuas and Real, respectively, whereas the EC50 was at a salinity of 30 ds m− 1 for Hualhuas plants and at 10–20 ds m− 1 for Real plants. Taken together, the relative decline in biomass, salinity resistance threshold, and EC50, indicate that the Peruvian cultivar Hualhuas is more salt-resistant compared to the Bolivian cultivar Real. In general, high salt resistance requires a root system that is capable to sustain plant growth under stressful conditions. This was indeed the case for Hualhuas plants that responded by a general inhibition of the shoot growth, but with continued root growth, resulting in a reduction of the shoot-to-root fresh weight ratio from 13 to 9, compared to 13 to only 1 in Real plants. Several factors may act as a bottleneck for plant growth under high salinity [52]. The primary deleterious constraint of salinity on plant growth is due to an osmotic effect [53]. As shown in Fig. (3), ψs of all plant organs of both cultivars dropped gradually and became more negative as water salinity rose, the effect that was more obvious for Real plants. In accordance with other studies, this implies that quinoa has a very efficient system to adjust osmotically and preserve a positive water balance under saline conditions [16]. This behavior is reflected by the trends of improved water content in all plant organs of both cultivars, particularly under moderate salinities (Fig. 2A and B).

However, high water salinity reduced the water contents of all plant organs of both cultivars, with less severe effects on Hualhuas plants. Against these findings and in accordance with Eisa et al. [16], one can presume that osmotic constraint is not a limiting factor for the growth of both cultivars under saline conditions. For both cultivars, salt-induced reduction in ψs was associated concurrently with substantial Na+ accumulation in all plant organs, particularly in the shoots. This confirms that both cultivars behave as salt-includers, utilizing controlled uptake of inorganic ions to adjust osmotically [16, 19]. Osmotic adjustment by massive accumulation of inorganic ions has been amply reported in many halophytic species, including quinoa [16, 34, 52]. Yet, NaCl salinity progressively increased Na+ contents, but significantly decreased those of K+ in all plant organs for both quinoa cultivars. This consequently led to a general trend of salt-induced reduction in K+/Na+ ratio, although with significant differences between cultivars (Fig. 4A and B). This might be due to a competition between K+ and Na+ uptake or due to changes in membrane integrity caused by the displacement of Ca2+ by Na+ under saline conditions [54]. In this study, significant differences in Na+ accumulation were observed between quinoa cultivars. Plants of Hualhuas accumulated distinctly lower amount of Na+ in all organs as compared with Real plants. This would explain, at least in part, the lower (more negative values) ψs observed in the tissues of this cultivar compared to Hualhuas (Fig. 3A and B). Another possible explanation for the differences in Na+ accumulation could be due to genotypic differences in the rates of Na+ uptake and xylem loading between these cultivars, which remains to be elucidated. Salt accumulation in excess of what is required for osmotic adjustment may lead to tissue dehydration, ion imbalance, and/or ion toxicity. Such conditions may result in growth reduction and inhibition of new leaf initiation, and consequently lower salt resistance as observed for Real plants. Hence, it is plausible to suggest that Hualhuas plants exert more efficient mechanisms to control Na+ uptake, translocation, and sequestration at the whole plant level. This is supported by the results of Shabala et al. [11], who found a strong negative correlation between xylem Na+ content and salinity resistance and suggested that controlling xylem Na+ loading is more important than root Na+ exclusion from uptake for salinity resistance in quinoa. At the cellular level, Na+ sequestration into the vacuoles is crucial to avoid its toxic effects in the cytosol, while providing a cheap osmoticum for water uptake and turgor maintenance [55]. Nonetheless, ion sequestration by itself is an energy-consuming process (Na+ accumulation occurs against a concentration gradient) and is usually accompanied by a concurrent increase in cytosolic osmolality to counteract the high osmolality of vacuoles [52]. This would be achieved by either increasing cytosolic K+, or accumulating organic osmolytes (extra energy requirements) in this compartment. As the latter is an energetically expensive option that may cause growth reductions, K+ accumulation is much more preferred under high salinity [43, 56]. Reportedly, salinity resistance in quinoa is attributed to its highly efficient K+ retention [15, 34, 57]. In the present study, Hualhuas plants maintained a higher K+/Na+ ratio in their organs compared to Real plants (Fig. 4). In agreement with Eisa et al. [16], maintaining adequate K+/Na+ in the tissues of Hualhuas plants, especially in the roots and juvenile leaves could be considered as advantageous, insofar, because it means more K+ is directed to the most actively growing tissues (where metabolic demands are expected to be greatest and Na+ sensitivity is highest). Increases in organic osmolytes have been also reported in quinoa [16, 24, 39]. This is also shown by our data: proline concentration increased by 40–53% in all organs of Hualhuas plants (more salt-resistant) in response to salinity (Fig. 5A). The same trend was observed for Real (less salt-resistant), but only in the juvenile leaves (75% increases). On contrary, proline concentration in the roots and adult leaves were reduced by 12% and 47%, respectively (Fig. 5B). Ruiz-Carrasco et al. [39] showed that 300 mM NaCl induced proline accumulation in all quinoa accessions tested, the effect that was more pronounced for the most salt-resistant genotypes.

Inadequate cytosolic ion ratios (K+/Na+, Mg2+/Na+, and Ca2+/Na+) can impair the plant photosynthetic machinery. Data of the present study showed that the concentrations of Chl(a) and Chl(b) were gradually and significantly (P < 0.05) reduced in both quinoa cultivars with increasing water salinity (Fig. 6A and B). Reduction in chlorophyll concentrations has been widely reported under saline conditions [16, 37] and attributed mainly to ion deficiency, disturbance of chloroplast membranes, instability of protein complexes and enhanced chlorophyllase activity [58, 59]. Our findings also clearly showed that salt-mediated reductions in Chl(a) and Chl(b) were more obvious for Hualhuas plants compared to Real (Fig. 6A and B). In agreement with Geissler et al. [47], reduction in chlorophyll contents seems to be an adaptive mechanism (from an energetic point of view) to cope with salt stress, since it may lead to reduce the over reduction of the photosynthetic electron transport chain and hence the generation of ROS. Nevertheless, this would lead to the decline of the photosynthetic capacity. As shown in Fig. (7) and Table (2), Anet was greatly decreased in both cultivars as water salinity rose, with more adverse effect on Hualhuas plants. Salinity stress led to reduce photosynthesis in a wide variety of halophytic species [37, 60] as well as in several quinoa cultivars [23, 45, 47]. Because plant growth and productivity are inextricably related to its photosynthetic capacity, inhibition of the latter as a response to salinity stress is suggested to be responsible, at least in part, for the reduction in plant growth and yield [52]. In both quinoa cultivars under the study, NaCl salinity affected Anet a priori by enhancing stomatal closure, (stomatal limitation), resulting into substantial reductions in CO2 diffusion to the carboxylation sites. This interpretation is supported by the linear proportionality between Anet and gs (Table 2). Similarly, a positive correlation between the photosynthetic rate and stomatal conductance has been noted in quinoa [10, 16, 47]. Salt-induced reductions in gs were higher in Hualhuas plants (93.5%) when compared to Real ones (77.9%) at full-strength salinity treatment (Table 2). Salt-induced inhibition in gs was accompanied by a progressive decline in E (Table 2), likely contributed to a positive water balance. Similar features for water conservation were also reported in quinoa [47] under saline conditions. Salt-induced reduction in E was higher in Hualhuas plants (Table 2), further suggesting that this cultivar is better adapted to high salinity. Lower E can represent an additional adaptive mechanism under high salinity, as it would reduce salt loading into the leaves and hence prolong leaf lifespan by maintaining a subtoxic level of salts [61]. This, indeed, may explain the lower Na+ accumulation and thus the higher K+/Na+ observed for Hualhuas plants. As mentioned above, reduced gs might limit the uptake and diffusion of CO2 to the carboxylation sites, as reflected by decreased Ci and hence Ci/Ca ratio (Table 2), thus resulting into an impaired Anet. The conspicuously low Ci concentration and the linear correlation between gs, E, Ci, and Anet in Hualhuas plants indicate that the limitation of photosynthesis under high salinity conditions in this cultivar is mainly a stomatal one (restricted by stomatal closure and substrate deficiency) [47]. However, this was not the case for Real plants, as high salinity led to a marked increase in Ci/Ca to reach the control values (Table 2). This suggests that stomatal closure is not a limiting factor for photosynthesis in Real under high saline conditions. Impaired photosynthesis in salt-stressed plants can be also due to the leaf biochemical and photochemical (non-stomatal) limitations [16]. This may be due to a decrease in Rubisco activity and/or content, a reduction in RuBP or Pi regeneration, or a decrease in PSII photochemistry efficiency [49]. The latter can decrease light absorption by the leaf [62], which is reflected by lower Ls in both cultivars under saline conditions, particularly, in Hualhuas plants (Fig. 7; Table 1). This led, in turn, to a significant reduction (optimization) in ΦCO2 in both cultivars (Table 1). Similar results have been previously reported and interpreted as an important mechanism to reduce the over-reduction of PSII and PSI and hence the generation of reactive oxygen species (ROS) [47]. Salinity-induced reduction in E was proportionally higher than that of Anet, leading to enhance PWUE by more than three folds in Hualhuas plants at the highest salinity treatment (Table 2). Such an increase in PWUE has been observed for many halophytic species, including quinoa in response to salinity stress [52, 63]. Consistent with previous studies [47, 64], salt-induced improvement in PWUE would be an advantage, conferring long-term survival of Hualhuas plants under stress conditions. The same trend of an enhanced PWUE was also observed for Real plants in response to salinity (Table 2), but only up to a salinity level of 400 mM NaCl, thereafter, PWUE was drastically decreased by 88.8% in these plants (Table 2). Apparently, this could explain the relatively low salt resistance of Real plants compared to Hualhuas ones.

Conclusion

Taken together, our results justified the potential of quinoa as a highly salt-resistant species (in terms of biomass production) able to grow even at 100% seawater salinity (sws). Both quinoa cultivars shared many common features of salt resistance mechanisms, although significant differences in their growth responses were observed. Osmotic constraint was not a major reason for the reduced growth in both cultivars under saline conditions. Rather, salt-induced growth reduction was presumably due to ion deficiency and/or toxicity, leading consequently to an impaired photosynthetic capacity. Results of this study allow for the speculation that the Peruvian cultivar Hualhuas is more salt resistant compared to the Bolivian cultivar Real. This might be largely attributed to a more efficient control mechanism on xylem Na+ loading and better K+ retention, ensuring a higher K+/Na+ ratio compared to Real plants. Its lower energy demand and higher responsiveness to balance photosynthesis may also contribute to its higher degree of salt resistance. Finally, it should be mentioned that the Peruvian cultivar Hualhuas is not only a promising candidate, suitable for the Egyptian conditions, but also through a deep understanding of its physiological and molecular resistance mechanisms, would provide a possible route to enhance salinity resistance in other genotypes.

Materials and methods

Plant materials, experimental design, and growth conditions

The present study was performed at the controlled greenhouse of the Agricultural Botany Department, Faculty of Agriculture, Ain Shams University, Qalyubia Governorate, Egypt (Latitude 30° 06′ 42″ N; Longitude 31° 14′ 46″ E), to investigate the eco-physiological responses of two quinoa cultivars grown under saline conditions. Seeds of C. quinoa cv. Hualhuas (origin: International Potato Center, CIP, Lima, Peru) and C. quinoa cv. Real (origin: Salar de Uyuni, Bolivia) were surface-sterilized with 70% ethanol for 1 min and subsequently with 0.5% NaOCl for 3 min before they were rinsed thoroughly with sterile water. The seeds were then sown in black plastic pots (30 cm diameter and 21 cm height), filled with washed sand (8 kg each, on a dry weight basis), five seeds per pot. The pots were kept on a bench at ambient temperatures of 22 ± 3 ºC daytime and 14 ± 3.5 ºC nighttime, a photoperiod of 10 h, relative humidity of 60–70%, and light intensity of 1500–2000 µmol m− 2 s− 1. The plants were irrigated manually and regularly with a nutrient solution [65]. After the emergence of the first two true leaves (three weeks after the germination), the plants were thinned to two seedlings of uniform size per pot. Salinity treatment started after a period of another two weeks by raising NaCl concentration in the nutrient solution in steps of 100 mM NaCl daily until the final concentrations were achieved to avoid salt shock injuries. There were altogether six salinity treatments (eight replicate pots for each treatment): control, 100, 200, 300, 400, and 500 mM NaCl [equivalent to 0, 20, 40, 60, 80, and 100% seawater salinity (sws)]. Salinity treatments were performed for eight weeks.

Harvest procedure and growth parameter measurements

The plants were destructively harvested eight weeks after the initiation of salinity treatment (six replicates each treatment). They were separated into roots (R), stems (S), adult leaves (La), juvenile leaves (Lj), and inflorescences (In). The root segments were gently cleaned from sand, washed for 1–2 min with ice-cold distilled water to remove the excess nutrients and salts, and then blotted carefully with tissue paper to remove adhered surface water. The fresh weights of all plant organs were directly recorded. Representative specimens of about 500–1000 mg from each plant organ (R, La, and Lj) were taken and stored at -20 ºC for further quantitative chemical analyses. To obtain the dry weights of different plant organs, specimens of about 500 mg were dried at 70 ºC until they reached a constant weight and the water content was determined as percentages of the fresh weights.

Determination of osmotic potential

The osmotic potential (ψs) of the press sap of R, La, and Lj was measured using the freeze-point depression method using an Osmometer (Osmomat 030, Genotec GMBH, Berlin). A 300 mOsmol NaCl solution was used as a standard and the readings were then converted to pressure units using a conversion table according to H-W Koyro [66].

Determination of mineral elements

Approximately 0.2 g of pulverized dried material from R, La, and Lj were weighed and wet digested using concentrated sulphuric acid (H2SO4) and hydrogen peroxide (H2O2 30%). The cleared, cooled extracts were carefully completed to a final volume of 50 ml with distilled water and then filtered through Whatman filter paper No. 42. Potassium (K+) and sodium (Na+) concentrations in these extracts were measured using a flame emission photometer method (JENWAY, PFP-7, ELE Instrument Co. Ltd., Essex, UK).

Photosynthetic pigments and gaseous exchange measurements

Determination of photosynthetic pigments

An appropriate amount of fresh materials (ten disks) from the uppermost fully expanded juvenile leaves were extracted in 80% (v/v) aqueous acetone. The concentrations of chlorophyll (a), chlorophyll (b), and carotenoids were determined spectrophotometrically as described by HK Lichtenthaler [67].

Leaf gas exchange measurements

The responses of leaf CO2/H2O gas exchange parameters to different water salinities and light intensities were assessed using an open portable photosynthesis measurement system (LI-COR 6400, Lincoln, NE, USA). One week before harvest, several photosynthetic parameters such as net assimilation rate (Anet, µmol CO2 m− 2s− 1), transpiration rate (E, mmol H2O m− 2s− 1), stomatal conductance (gs, mmol m− 2s− 1) and intercellular CO2 concentration (Ci, µmol mol− 1) were determined at various photosynthetic active radiation (PAR = 0, 400, 800, 1200, 1600 and 2000 µmol quanta m− 2 s− 1). These PAR values were provided with an artificial LED light source (6400-02B, LI-COR, Lincoln, NE, USA). The relative humidity was maintained at 50–60%, leaf temperature was set at 25 °C, the flow rate was set at 300 µmol s− 1, and CO2 concentration was maintained at 400 µmol mol− 1 inside the leaf chamber. All measurements were achieved between 09:00 and 15:00 o’clock. Assimilation parameters were recorded at each light level following an acclimation period of 5 min and measurements were repeated to obtain, at least six, stable readings for each salinity treatment. Photosynthesis water use efficiency (PWUE, defined as the ratio between net assimilation rate and transpiration) was calculated by the LI-6400xtdata analysis program (LI-COR, Lincoln, NE, USA). Values of Anet were plotted against PAR and fitted to A-PAR response curves with SigmaPlot 12.0 software (Systat Software, Inc.) using an exponential function as explained by Schulte et al. [68]. By means of this function, the initial linear slope of the light response curve, which describes the efficiency of photosynthetic energy conversion in leaves at sub-saturating light intensities (ΦCO2), the light compensation point (Lc, the value of PAR when Anet = 0), the light saturation point (Ls, the value of PAR when Anet= 90% Amax) and the dark respiration rate (Rd) were calculated.

Determination of proline

The proline contents of different plant organs (R, La, and Lj) were determined according to the method of Bates et al. [69]. The absorbance of the toluene phase was read using a UV/VIS spectrophotometer (T-60, PG instrument, Wibtoft Leicestershire, UK), at a wavelength of 520 nm, and proline concentration was calculated by comparing sample absorbencies with the standard proline curve.

Statistical analysis

All data sets were subjected to one-way-ANOVA analysis using the SPSS for Windows statistical data analysis package (SPSS Inc., 2002, release 16, Chicago, Illinois,, USA) in order to determine if significant differences were found among means. To meet all assumptions for ANOVA, data transformation was performed when the original data were not normally distributed. Duncan’s multiple range test was employed to determine if significant (P < 0.05) differences occurred between individual treatments.

Data Availability

The data sets generated during the current study are available from the first author on reasonable request.

Abbreviations

A net :

Net CO2 assimilation rate

C a :

Ambient CO2 concentration

C i :

Intercellular CO2 concentration

C i /C a :

Ratio of intercellular to ambient CO2 concentration

E :

Transpiration rate

EC50 :

Water salinity which reduces the maximum yield by 50%

g s :

Stomatal conductance

InFW:

Inflorescence fresh weight

L c :

Light compensation point

LaFW:

Adult leaves fresh weight

LjFW:

Juvenile leaves fresh weight

L s :

Saturation irradiance

PAR :

Photosynthetic active radiation

PFW:

Plant fresh weight

PWUE :

Photosynthetic water use efficiency

R d :

Dark respiration

RFW:

Root fresh weight

ROS:

Reactive oxygen species

SFW:

Stem fresh weight

sws:

Seawater salinity

Φ CO2 :

Apparent carboxylation efficiency

References

  1. Singh K. Microbial and enzyme activities of saline and sodic soils. Land Degrad Dev. 2016;27(3):706–18.

    Article  Google Scholar 

  2. Hajiboland R. Role of arbuscular mycorrhiza in amelioration of salinity. In: Salt Stress in Plants Springer; 2013: 301–354.

  3. Gupta B, Huang B. Mechanism of salinity tolerance in plants: physiological, biochemical, and molecular characterization. International journal of genomics 2014, 2014.

  4. Lavania D, Dhingra A, Siddiqui MH, Al-Whaibi MH, Grover A. Current status of the production of high temperature tolerant transgenic crops for cultivation in warmer climates. Plant Physiol Biochem. 2015;86:100–8.

    Article  CAS  PubMed  Google Scholar 

  5. Hasnain M, Munir N, Abideen Z, Zulfiqar F, Koyro HW, El-Naggar A, Caçador I, Duarte B, Rinklebe J, Yong JWH. Biochar-plant interaction and detoxification strategies under abiotic stresses for achieving agricultural resilience: a critical review. Ecotoxicol Environ Saf. 2023;249:114408.

    Article  CAS  PubMed  Google Scholar 

  6. Panagea I, Daliakopoulos I, Tsanis I, Schwilch G. Evaluation of promising technologies for soil salinity amelioration in Timpaki (Crete): a participatory approach. In: Solid Earth vol. 7; 2016: 177.

  7. González JA, Hinojosa L, Mercado MI, Fernández-Turiel J-L, Bazile D, Ponessa GI, Eisa S, González DA, Rejas M, Hussin S. A long journey of CICA-17 quinoa variety to salinity conditions in Egypt: Mineral concentration in the seeds. Plants. 2021;10(2):407.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Rozema J, Muscolo A, Flowers T. Sustainable cultivation and exploitation of halophyte crops in a salinising world. Environ Exp Bot. 2013;92:1–3.

    Article  Google Scholar 

  9. Tarek K, Alshamy I, Hussin S, Eisa S. Nitrogen Nutrition and Adaptation of Halophyte Chenopodium quinoa to salt stress. Arab Universities Journal of Agricultural Sciences. 2021;29(2):491–503.

    Google Scholar 

  10. Adolf VI, Shabala S, Andersen MN, Razzaghi F, Jacobsen S-E. Varietal differences of quinoa’s tolerance to saline conditions. Plant Soil. 2012;357(1–2):117–29.

    Article  CAS  Google Scholar 

  11. Shabala S, Hariadi Y, Jacobsen S-E. Genotypic difference in salinity tolerance in quinoa is determined by differential control of xylem Na+ loading and stomatal density. J Plant Physiol. 2013;170(10):906–14.

    Article  CAS  PubMed  Google Scholar 

  12. Bazile D, Jacobsen S-E, Verniau A. The global expansion of quinoa: trends and limits. Front Plant Sci. 2016;7:622.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Fuentes F, Bazile D, Bhargava A, Martinez EA. Implications of farmers’ seed exchanges for on-farm conservation of quinoa, as revealed by its genetic diversity in Chile. J Agricultural Sci. 2012;150(6):702–16.

    Article  Google Scholar 

  14. Tapia ME. the Long Journey of quinoa: Who wrote its history? STATE OF THE ART REPORT ON QUINOA 2015:3.

  15. Cai Z-Q, Gao Q. Comparative physiological and biochemical mechanisms of salt tolerance in five contrasting highland quinoa cultivars. BMC Plant Biol. 2020;20(1):70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Eisa S, Hussin S, Geissler N, Koyro H. Effect of NaCl salinity on water relations, photosynthesis and chemical composition of Quinoa (Chenopodium quinoa Willd.) As a potential cash crop halophyte. Aust J Crop Sci. 2012;6(2):357.

    CAS  Google Scholar 

  17. Razzaghi F, Jacobsen S-E, Jensen CR, Andersen MN. Ionic and photosynthetic homeostasis in quinoa challenged by salinity and drought–mechanisms of tolerance. Funct Plant Biol. 2015;42(2):136–48.

    Article  CAS  PubMed  Google Scholar 

  18. Pulvento C, Lavini A, Riccardi M, d’Andria R, Ragab R. Assessing amaranth adaptability in a Mediterranean area of south Italy under different climatic scenarios. Irrig Sci. 2015;64(1):50–8.

    Google Scholar 

  19. Panuccio M, Jacobsen S-E, Akhtar SS, Muscolo A. Effect of saline water on seed germination and early seedling growth of the halophyte quinoa. AoB plants. 2014;6:plu047.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Adolf VI, Jacobsen S-E, Shabala S. Salt tolerance mechanisms in quinoa (Chenopodium quinoa Willd). Environ Exp Bot. 2013;92:43–54.

    Article  CAS  Google Scholar 

  21. González JA, Gallardo M, Hilal M, Rosa M, Prado FE. Physiological responses of quinoa (Chenopodium quinoa Willd.) To drought and waterlogging stresses: dry matter partitioning. Bot Stud. 2009;50(1):35–42.

    Google Scholar 

  22. Hussin S, Khalifa W, Geissler N, Koyro HW. Influence of the root endophyte Piriformospora indica on the plant water relations, gas exchange and growth of Chenopodium quinoa at limited water availability. J Agron Crop Sci. 2017;203(5):373–84.

    Article  CAS  Google Scholar 

  23. Razzaghi F, Ahmadi SH, Adolf VI, Jensen CR, Jacobsen SE, Andersen MN. Water relations and transpiration of quinoa (Chenopodium quinoa Willd.) Under salinity and soil drying. J Agron Crop Sci. 2011;197(5):348–60.

    Article  Google Scholar 

  24. Jacobsen S-E, Monteros C, Corcuera LJ, Bravo LA, Christiansen JL, Mujica A. Frost resistance mechanisms in quinoa (Chenopodium quinoa Willd). Eur J Agron. 2007;26(4):471–5.

    Article  Google Scholar 

  25. González JA, Rosa M, Parrado MF, Hilal M, Prado FE. Morphological and physiological responses of two varieties of a highland species (Chenopodium quinoa Willd.) Growing under near-ambient and strongly reduced solar UV–B in a lowland location. J Photochem Photobiol B. 2009;96(2):144–51.

    Article  PubMed  Google Scholar 

  26. Bertero H, De la Vega A, Correa G, Jacobsen S, Mujica A. Genotype and genotype-by-environment interaction effects for grain yield and grain size of quinoa (Chenopodium quinoa Willd.) As revealed by pattern analysis of international multi-environment trials. Field Crops Research. 2004;89(2–3):299–318.

    Article  Google Scholar 

  27. El-Sayed E-SE-S, Abd El-Samad EE-DH, Ahmed HAA, AbouSekken MSM. Enhancement of Quinoa Grain Yield and Nutritional Quality by Potassium Fertilization Combined with Foliar Spraying of Seaweed Extract. J Ecol Eng. 2023;24(3):341–56.

    Article  Google Scholar 

  28. Repo-Carrasco R, Espinoza C, Jacobsen S-E. Nutritional value and use of the Andean crops quinoa (Chenopodium quinoa) and kañiwa (Chenopodium pallidicaule). Food reviews international. 2003;19(1–2):179–89.

    Article  Google Scholar 

  29. Gordillo-Bastidas E, Díaz-Rizzolo D, Roura E, Massanés T, Gomis R. Quinoa (Chenopodium quinoa Willd), from nutritional value to potential health benefits: an integrative review. J Nutr Food Sci. 2016;6(497):104172.

    Google Scholar 

  30. Ruiz KB, Biondi S, Oses R, Acuña-Rodríguez IS, Antognoni F, Martinez-Mosqueira EA, Coulibaly A, Canahua-Murillo A, Pinto M, Zurita-Silva A. Quinoa biodiversity and sustainability for food security under climate change. A review. Agron Sustain Dev. 2014;34(2):349–59.

    Article  Google Scholar 

  31. Stikic R, Glamoclija D, Demin M, Vucelic-Radovic B, Jovanovic Z, Milojkovic-Opsenica D, Jacobsen S-E, Milovanovic M. Agronomical and nutritional evaluation of quinoa seeds (Chenopodium quinoa Willd.) As an ingredient in bread formulations. J Cereal Sci. 2012;55(2):132–8.

    Article  CAS  Google Scholar 

  32. Small E. Quinoa: is the United Nations’ featured crop of 2013 bad for biodiversity. Biodiversity. 2013;14(3):169–79.

    Article  Google Scholar 

  33. Prado FE, Boero C, Gallardo M, Gonzalez JA. Effect of NaCl on germination, growth, and soluble sugar content in Chenopodium quinoa Willd. Seeds. Bot Bull Acad Sinica 2000, 41.

  34. Hariadi Y, Marandon K, Tian Y, Jacobsen S-E, Shabala S. Ionic and osmotic relations in quinoa (Chenopodium quinoa Willd.) Plants grown at various salinity levels. J Exp Bot. 2011;62(1):185–93.

    Article  CAS  PubMed  Google Scholar 

  35. Flowers TJ, Munns R, Colmer TD. Sodium chloride toxicity and the cellular basis of salt tolerance in halophytes. Ann Botany. 2015;115(3):419–31.

    Article  CAS  Google Scholar 

  36. Hussin S, Geissler N, El-Far MM, Koyro H-W. Effects of salinity and short-term elevated atmospheric CO2 on the chemical equilibrium between CO2 fixation and photosynthetic electron transport of Stevia rebaudiana Bertoni. Plant Physiol Biochem. 2017;118:178–86.

    Article  CAS  PubMed  Google Scholar 

  37. Geissler N, Hussin S, Koyro H-W. Interactive effects of NaCl salinity and elevated atmospheric CO2 concentration on growth, photosynthesis, water relations and chemical composition of the potential cash crop halophyte Aster tripolium L. Environ Exp Bot. 2009;65(2–3):220–31.

    Article  CAS  Google Scholar 

  38. Gómez-Pando LR, Álvarez‐Castro R, Eguiluz‐de la Barra A. Effect of salt stress on peruvian germplasm of Chenopodium quinoa Willd.: a promising crop. J Agron Crop Sci. 2010;196(5):391–6.

    Article  Google Scholar 

  39. Ruiz-Carrasco K, Antognoni F, Coulibaly AK, Lizardi S, Covarrubias A, Martínez EA, Molina-Montenegro MA, Biondi S, Zurita-Silva A. Variation in salinity tolerance of four lowland genotypes of quinoa (Chenopodium quinoa Willd.) As assessed by growth, physiological traits, and sodium transporter gene expression. Plant Physiol Biochem. 2011;49(11):1333–41.

    Article  CAS  PubMed  Google Scholar 

  40. Orsini F, Accorsi M, Gianquinto G, Dinelli G, Antognoni F, Carrasco KBR, Martinez EA, Alnayef M, Marotti I, Bosi S. Beyond the ionic and osmotic response to salinity in Chenopodium quinoa: functional elements of successful halophytism. Funct Plant Biol. 2011;38(10):818–31.

    Article  CAS  PubMed  Google Scholar 

  41. Peterson A, Murphy K. Tolerance of lowland quinoa cultivars to sodium chloride and sodium sulfate salinity. Crop Sci. 2015;55(1):331–8.

    Article  CAS  Google Scholar 

  42. Ruiz K, Biondi S, Martínez E, Orsini F, Antognoni F, Jacobsen S-E. Quinoa–a model crop for understanding salt-tolerance mechanisms in halophytes. Plant Biosystems-An International Journal Dealing with all Aspects of Plant Biology. 2016;150(2):357–71.

    Article  Google Scholar 

  43. Huchzermeyer B, Koyro H. Salt and drought stress effects on photosynthesis. Handbook of plant and crop stress, 2nd edn Marcel Dekker Inc, New York 2005:751–78.

  44. Sanchez HB, Lemeur R, Damme Pv, Jacobsen S-E. Ecophysiological analysis of drought and salinity stress of quinoa (Chenopodium quinoa Willd). Food Reviews International. 2003;19(1–2):111–9.

    Article  Google Scholar 

  45. González JA, Eisa SS, Hussin SA, Prado FE. Quinoa: an incan crop to face global changes in agriculture. Quinoa: Improv sustainable Prod 2015:1–18.

  46. Atkin O. New Phytologist and the ‘fate’of carbon in terrestrial ecosystems. New Phytol. 2015;205(1):1–3.

    Article  PubMed  Google Scholar 

  47. Geissler N, Hussin S, El-Far MM, Koyro H-W. Elevated atmospheric CO2 concentration leads to different salt resistance mechanisms in a C3 (Chenopodium quinoa) and a C4 (Atriplex nummularia) halophyte. Environ Exp Bot. 2015;118:67–77.

    Article  CAS  Google Scholar 

  48. Shabala L, Mackay A, Tian Y, Jacobsen SE, Zhou D, Shabala S. Oxidative stress protection and stomatal patterning as components of salinity tolerance mechanism in quinoa (Chenopodium quinoa). Physiol Plant. 2012;146(1):26–38.

    Article  CAS  PubMed  Google Scholar 

  49. Lawlor DW, Tezara W. Causes of decreased photosynthetic rate and metabolic capacity in water-deficient leaf cells: a critical evaluation of mechanisms and integration of processes. Ann Botany. 2009;103(4):561–79.

    Article  CAS  Google Scholar 

  50. Shannon M, Grieve C. Tolerance of vegetable crops to salinity. Sci Hort. 1998;78(1–4):5–38.

    Article  Google Scholar 

  51. Del Castillo C, Winkel T, Mahy G, Bizoux J-P. Genetic structure of quinoa (Chenopodium quinoa Willd.) From the bolivian altiplano as revealed by RAPD markers. Genet Resour Crop Evol. 2007;54(4):897–905.

    Article  Google Scholar 

  52. Hussin S, Geissler N, Koyro H-W. Effect of NaCl salinity on Atriplex nummularia (L.) with special emphasis on carbon and nitrogen metabolism. Acta Physiol Plant. 2013;35(4):1025–38.

    Article  CAS  Google Scholar 

  53. Munns R. Genes and salt tolerance: bringing them together. New Phytol. 2005;167(3):645–63.

    Article  CAS  PubMed  Google Scholar 

  54. Silveira JAG, Araújo SAM, Lima JPMS, Viégas RA. Roots and leaves display contrasting osmotic adjustment mechanisms in response to NaCl-salinity in Atriplex nummularia. Environ Exp Bot. 2009;66(1):1–8.

    Article  CAS  Google Scholar 

  55. Shabala S, Mackay A. Ion transport in halophytes. Advances in botanical research. Vol. 57: Elsevier; 2011: 151–99.

  56. Chen Z, Zhou M, Newman IA, Mendham NJ, Zhang G, Shabala S. Potassium and sodium relations in salinised barley tissues as a basis of differential salt tolerance. Funct Plant Biol. 2007;34(2):150–62.

    Article  CAS  PubMed  Google Scholar 

  57. Shabala S, Shabala S, Cuin TA, Pang J, Percey W, Chen Z, Conn S, Eing C, Wegner LH. Xylem ionic relations and salinity tolerance in barley. Plant J. 2010;61(5):839–53.

    Article  CAS  PubMed  Google Scholar 

  58. Fang Z, Bouwkamp JC, Solomos T. Chlorophyllase activities and chlorophyll degradation during leaf senescence in non-yellowing mutant and wild type of Phaseolus vulgaris L. J Exp Bot. 1998;49(320):503–10.

    CAS  Google Scholar 

  59. Santos CV. Regulation of chlorophyll biosynthesis and degradation by salt stress in sunflower leaves. Sci Hort. 2004;103(1):93–9.

    Article  CAS  Google Scholar 

  60. Koyro H-W, Hussain T, Huchzermeyer B, Khan MA. Photosynthetic and growth responses of a perennial halophytic grass Panicum turgidum to increasing NaCl concentrations. Environ Exp Bot. 2013;91:22–9.

    Article  CAS  Google Scholar 

  61. Everard JD, Gucci R, Kann SC, Flore JA, Loescher WH. Gas exchange and carbon partitioning in the leaves of celery (Apium graveolens L.) at various levels of root zone salinity. Plant Physiol. 1994;106(1):281–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Christian R. Interactive effects of salinity and irradiance on photoprotection in acclimated seedlings of two sympatric mangroves. Trees. 2005;19(5):596–606.

    Article  CAS  Google Scholar 

  63. Redondo-Gómez S, Mateos-Naranjo E, Davy AJ, Fernández-Muñoz F, Castellanos EM, Luque T, Figueroa ME. Growth and photosynthetic responses to salinity of the salt-marsh shrub Atriplex portulacoides. Ann Botany. 2007;100(3):555–63.

    Article  Google Scholar 

  64. Gleick PH, Christian-Smith J, Cooley H. Water-use efficiency and productivity: rethinking the basin approach. Water Int. 2011;36(7):784–98.

    Article  Google Scholar 

  65. Hoagland DR, Arnon DI. The water-culture method for growing plants without soil. Circular California agricultural experiment station 1950, 347(2nd edit).

  66. Koyro H-W. Study of potential cash crop halophytes by a quick check system: determination of the threshold of salinity tolerance and the ecophysiological demands. Cash crop halophytes: recent studies. Springer; 2003: 5–17.

  67. Lichtenthaler HK. Chlorophylls and carotenoids: pigments of photosynthetic biomembranes. Methods in enzymology. Vol. 148: Elsevier; 1987: 350–82.

  68. Schulte M, Offer C, Hansen U. Induction of CO2-gas exchange and electron transport: comparison of dynamic and steady-state responses in Fagus sylvatica leaves. Trees. 2003;17(2):153–63.

    Article  CAS  Google Scholar 

  69. Bates L, Waldren R, Teare I. Rapid determination of free proline for water-stress studies. Plant Soil. 1973;39(1):205–7.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The authors are very grateful for the financial support from the Science, Technology & Innovation Funding Authority (STDF), Academy of Scientific Research and Technology (ASRT), Egypt, and the Saline Agricultural Unit, Faculty of Agriculture, Ain Shams University, Egypt.

Funding

This work was partially supported by the Science, Technology & Innovation Funding Authority (STDF), Academy of Scientific Research and Technology (ASRT), Egypt, through an agricultural target research project “Quinoa as a new non-traditional grain, oil, vegetable, and fodder crop in Egypt, ID-2552”.

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB).

Author information

Authors and Affiliations

Authors

Contributions

Hussin, Eisa, and Ali designed and supervised this study. Lotfy, Hussin, and Eisa performed experiments. Hussin, Lotfy, Eid, and Abd El-Samad estimated all biochemical and physiological analyses. Hussin performed data interpretation and wrote the original draft, software, and statistical analysis. Hussin, Eisa, and Abd-Elkader revised the manuscript grammatically. All authors have read and approved the manuscript.

Corresponding authors

Correspondence to Sayed A. Hussin or Sayed Said Eisa.

Ethics declarations

Ethics approval and consent to participate

All plant-related procedures complied with the relevant institutional, national, and international guidelines and legislation.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hussin, S.A., Ali, S.H., Lotfy, M.E. et al. Morpho-physiological mechanisms of two different quinoa ecotypes to resist salt stress. BMC Plant Biol 23, 374 (2023). https://doi.org/10.1186/s12870-023-04342-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-023-04342-4

Keywords