Next Article in Journal
Analysis of Potato Physiological and Molecular Adaptation in Response to Different Water and Nitrogen Combined Regimes
Next Article in Special Issue
Membrane Fluidization Governs the Coordinated Heat-Inducible Expression of Nucleus- and Plastid Genome-Encoded Heat Shock Protein 70 Genes in the Marine Red Alga Neopyropia yezoensis
Previous Article in Journal
Effect of Green Synthesized ZnO-NPs on Growth, Antioxidant System Response and Bioactive Compound Accumulation in Echinops macrochaetus, a Potential Medicinal Plant, and Assessment of Genome Size (2C DNA Content)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization of Chlorophyll Fluorescence and Antioxidant Defense Parameters of Two Gracilariopsis lemaneiformis Strains under Different Temperatures

1
CAS and Shandong Province Key Laboratory of Experimental Marine Biology, Center for Ocean Mega-Science, Institute of Oceanology, Chinese Academy of Sciences, Qingdao 266071, China
2
Laboratory for Marine Biology and Biotechnology, Pilot National Laboratory for Marine Science and Technology (Qingdao), Qingdao 266071, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Plants 2023, 12(8), 1670; https://doi.org/10.3390/plants12081670
Submission received: 25 February 2023 / Revised: 5 April 2023 / Accepted: 6 April 2023 / Published: 17 April 2023
(This article belongs to the Special Issue Current Topics in Macroalgal Research)

Abstract

:
In this study, two Gracilariopsis lemaneiformis strains—the wild type and a green-pigmented mutant—were cultured at three temperatures (8, 20, and 30 °C) for 7 days to explore their temperature tolerance using photosynthetic performance and antioxidant defense parameters. When the two strains of G. lemaneiformis were separately cultured at 30 °C, the fast chlorophyll fluorescence intensity of the wild type decreased, whereas the green mutant showed no significant change. The decrease in the performance index on absorption basis value under heat stress was lower in the green mutant than in the wild type. In addition, the green mutant had stronger antioxidant activity at 30 °C. Furthermore, a greater decrease in the values of maximum photochemical quantum yield and performance index on an absorption basis in the green mutant indicated that it had a greater degree of inhibition of photosynthetic performance under low temperatures. However, the green mutant produced less reactive oxygen species under low temperatures, suggesting that the antioxidant potential of the green mutant might be higher. In conclusion, the green mutant exhibited heat tolerance and could recover from low-temperature damage; therefore, it has the potential for large-scale cultivation.

1. Introduction

Gracilariopsis lemaneiformis, an economically important seaweed, has been widely cultivated in China for over 20 years [1]. According to the National Bureau of Statistics [2], the annual yield of Gracilariales (mainly G. lemaneiformis) is the second largest in China. Besides being the main source of agar extraction [3], the alga has been widely used as fodder for abalone aquaculture and is ideal for carbon assimilation to remediate eutrophic ecosystems [4,5,6]. In addition, polysaccharides and their derivatives obtained from G. lemaneiformis have been found to perform various biological activities such as prebiotic, antiradiation, antitumor, immunomodulatory, and antioxidant activities [7,8,9].
Temperature is one of the key factors affecting the physiology and distribution of seaweeds [10]. Temperature can affect agar yield and biochemical components of G. lemaneiformis [11]. The suitable temperature for the growth of the wild population of G. lemaneiformis was between 12 °C and 23 °C [12]. Temperatures above 23 °C can disrupt the metabolic processes of G. lemaneiformis or even hinder its normal growth and development, leading to disease, death, and yield reduction [13,14]. The industrial production of G. lemaneiformis is limited by high temperatures in summer, especially along the coast of the South China Sea [15]. The heat-tolerant cultivars 981 and 2007, which can tolerate temperatures ranging from 23 °C to 26 °C, are widely cultivated in the South China Sea coastal region [3,16]. However, these cultivars still cannot survive in summer. With the increase in global warming, the temperature tolerance of G. lemaneiformis needs to be explored to improve its germplasm for aquaculture. Besides temperature, other factors, such as light and nutrient supply, are important for the growth of G. lemaneiformis [17].
Photosynthesis is one of the most sensitive physiological responses of plants, including seaweeds, to environmental changes. In the photosynthetic electron-transfer reaction, light energy fuels the electron transport from water to NADP+ via photosystem II (PSII), cytochrome b6f (Cyt b6f) complex, and photosystem I (PSI) to generate NADPH [18]. PSII is a multisubunit pigment transmembrane protein complex on the thylakoid membrane, consisting of D1, D2, CP43, and CP47 proteins [19]. PSII is sensitive to a wide range of environmental conditions and susceptible to damage; especially, the D1 protein is vulnerable to attack by reactive oxygen species (ROS) generated by reaction centers (RCs) [20]. Chlorophyll fluorescence is a non-invasive, reliable, and effective technique to assess the changes in photosynthetic performance, especially under stress conditions [21,22]. Moreover, chlorophyll fluorescence parameters are widely used to assess photosynthetic features in algae [23,24,25]. They have been used to study the stresses of temperature, carbon dioxide, light, and nutrient influence in G. lemaneiformis and to estimate the degree of stress on its photosynthetic apparatus under environmental conditions [26]. Using polyphasic chlorophyll fluorescence transients (OJIP) analysis, it was found that G. lemaneiformis was a more temperature-dependent than nutrient-dependent species [27]. The significant decrease in FV/Fm of G. lemaneiformis at an irradiance of 20 μmol photons m−2 s−1 (low light) indicated a decrease in its photosynthetic efficiency, which to some extent explained the mechanism of light irradiance stress inhibiting the growth of G. lemaneiformis [28]. Therefore, chlorophyll fluorescence parameters form a sensitive index for reflecting marine environmental influences on seaweed physiological processes.
It is known from various marine organisms that environmental perturbations, such as cold, heat, and oxidative stress, lead to the formation of ROS [29]. A moderate amount of ROS acts as a signal to induce an increase in antioxidant capacity, thereby increasing cellular stress resistance and viability [30]. However, excess amounts of ROS in seaweeds can cause damage to DNA, proteins, and membrane lipids, especially to the photosynthetic apparatus [31,32,33]. ROS was rapidly generated when G. lemaneiformis was exposed to heat stress [15]. Enzymatic antioxidant defenses, such as superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT), can scavenge ROS in seaweeds [6,28].
In the large-scale cultivation of G. lemaneiformis, excellent cultivar and germplasm are important factors to ensure its quality and yield, and the selection and promotion of G. lemaneiformis is a prerequisite for large-scale farming [34]. The acquisition of mutants helps in the search for superior germplasm. The G. lemaneiformis cultivar 981 was an excellent strain obtained from culture trials after nearly 20 years of domestication and screening of wild populations of G. lemaneiformis, which had the characteristics of high-temperature tolerance and high growth rate. The G. lemaneiformis cultivar 07-2 was obtained from the cultivar 981 by using N-methyl-N′-nitro-N-ni-trosoguanidine (MNNG), which has excellent characteristics of rapid growth and greater resistance to harsh environments [35]. The G. lemaneiformis cultivar Lulong No.1 is an excellent cultivar obtained by population selection, single cross-breeding, and UV mutagenic high-temperature stress screening using wild populations of G. lemaneiformis as the parent [36]. Huang et al. [37] analyzed the gene related to optically active phycoerythrin synthesis in both wild-type and green-pigmented mutant, and they found a reduction in optically active phycoerythrin in the green mutant. However, stress tolerance analyses of the G. lemaneiformis wild type and mutant have not been reported.
In the present study, the physiological responses of a green-pigmented mutant and the wild type of G. lemaneiformis to heat and cold stresses were analyzed using chlorophyll fluorescence and antioxidant defense parameters. The objectives of this study were (1) to compare the temperature tolerance between the wild type and the green mutant and (2) to provide a research base for the selection of high-quality G. lemaneiformis germplasm resources.

2. Results

2.1. Molecular Phylogenetic Analysis

The stable green-pigmented mutant (Figure 1a) was derived from the G. lemaneiformis cultivar 981 (Figure 1b) by spontaneous mutation and continuous selection for 3 years from 2019. Moreover, the molecular phylogenetic analysis (Figure 2) indicated that the green mutant and the wild type grouped with referenced G. lemaneiformis, forming a unique clade.

2.2. OJIP Fluorescence Transient Analysis

The OJIP transient curves of the two strains exhibited a polyphasic rise of O (20 μs), J (2 ms), I (30 ms), and P (maximum fluorescence) (Figure 3). After heat stress, the OJIP transient of the green mutant was similar to that at room temperature, indicating a better adaptation of the green mutant to heat stress (Figure 3a). However, all steps of the OJIP transient curve of the wild type showed a downward trend for heat stress conditions compared with normal temperature conditions (Figure 3b). In both strains, all steps of the OJIP transient curves were much lower for low-temperature conditions than for normal-temperature conditions. The lower fluorescence intensity of the OJIP transient in the two strains under low temperatures indicated an inhibition of photosynthetic performance.

2.3. Temperature Influences on Photosynthetic Electron Transport

In the green mutant, damage to the oxygen-evolving complex (WK), quantum yield for electron transport (φEo), ABS/RC, TRo/RC, ETo/RC, DIo/RC, and performance index (PIabs) values exhibited a downward trend, whereas maximum photochemical quantum yield (FV/Fm), the quantum yield of reduction of end electron acceptors of PSI (φRo), and RC/CSo values exhibited an upward trend after exposure to heat stress compared with normal temperature conditions (Figure 4a, Table S1).
In the wild type of G. lemaneiformis, FV/Fm, φEo, φRo, ETo/RC, RC/CSo, and PIabs values exhibited a downward trend, whereas WK, ABS/RC, and DIo/RC values exhibited an upward trend under heat stress compared with normal temperature conditions (Figure 4b). The increase in WK in the wild type indicated that the oxygen-evolving complex (OEC) might be damaged.
PIabs values of the two strains decreased after being exposed to heat stress (p < 0.05). Moreover, the PIabs value of the green mutant dropped by 16%, and that of the wild type decreased by 37% (p < 0.05), demonstrating that the PSII overall performance was less affected by heat stress in the green mutant than in the wild type. Therefore, high-temperature tolerance may be higher in the green mutant than in the wild type.
In the green mutant, WK, FV/Fm, φEo, φRo, TRo/RC, ETo/RC, RC/CSo, and PIabs values exhibited a downward trend, whereas ABS/RC and DIo/RC values exhibited an upward trend after exposure to low temperature compared with normal temperature conditions. In the wild type, FV/Fm, φEo, φRo, RC/CSo, and PIabs values exhibited a downward trend, whereas WK, ABS/RC, TRo/RC, ETo/RC, and DIo/RC values exhibited an upward trend under low temperature compared with normal-temperature conditions. Under low temperature, the FV/Fm value of the green mutant decreased by 22%, and that of the wild type decreased by 14%. The PIabs value decreased by 40% in the green mutant and 24% in the wild type (p < 0.05). The greater decrease in the FV/Fm and PIabs values in the green mutant indicated that it had a greater degree of inhibition of photosynthetic performance.

2.4. Non-Photochemical Quenching Detection

With exposure to heat stress, the non-photochemical quenching (NPQ) value increased three-fold (from 0.19 to 0.57) in the green mutant and nearly two-fold (from 0.42 to 0.73) in the wild type (Figure 5). The lower increase in the NPQ value in the wild type indicated a greater suppression of the photo-protection of NPQ. Interestingly, the NPQ value of the wild type under heat stress was 1.3-fold that of the green mutant. In addition, NPQ values in both strains were lower under low temperatures (Figure 5).

2.5. Oxygen-Free Radical and Lipid Peroxidation Measurements

OFR and MDA (lipid peroxidation level) content of the two strains were significantly high after heat stress compared with normal temperature conditions (Figure 6, p < 0.01). The OFR content in the green mutant and wild type increased by 21% and 62%, respectively, indicating that the wild type produced more ROS under heat stress. Moreover, the MDA content in the green mutant and wild type increased by 29% and 80%, respectively, showing that the green mutant had less lipid peroxidation, whereas the wild type had severe lipid peroxidation under heat stress.
The OFR content did not significantly change in the green mutant, but it significantly increased in the wild type after low-temperature treatment (Figure 6a, p < 0.01). The MDA content increased by 28% in the green mutant and 80% in the wild type (Figure 6b, p < 0.01), indicating more lipid peroxidation in the wild type under low temperature.

2.6. Antioxidant Enzyme Activity Measurements

The SOD, POD, and CAT activities of the two G. lemaneiformis strains increased under heat stress (Figure 7). The increased SOD (p < 0.01), POD (p < 0.01), and CAT (p < 0.01) activities in the green mutant indicate that it had a more efficient antioxidant system to protect from oxidative stress compared with the wild type. The low temperature did not significantly change POD and CAT activities in the two strains. There was a non-significant upward trend in SOD activity (p > 0.05) in the green mutant after low-temperature stress, while there was a weak increase in SOD activity (p < 0.01) in the wild type.

3. Discussion

3.1. Heat Stress in G. lemaneiformis

High summer temperatures can lead to algal decay, which limits the large-scale cultivation of G. lemaneiformis in China [38]. The two G. lemaneiformis strains showed different responses to heat stress. In plants and algae, PSII is the most sensitive component of the electron transport chain of photosynthesis in response to environmental changes [39]. OJIP fluorescence transient reflects the changes in the primary photochemical reaction of PSII and the electron transfer state of the photosynthetic structure [22]. The fluorescence intensity of the OJIP transient decreased in the G. lemaneiformis wild type at 30 °C, showing that high temperature may inhibit electron transfer on the PS II donor side or the PS II acceptor side. However, no significant effect was observed in the green mutant, indicating that the green mutant showed acclimation of photosynthesis under heat stress. To further investigate the acclimation of photosynthesis under heat stress in the green mutant, we calculated several JIP parameters that identify sensitive functions such as energy absorption, energy capture, and electron transport in the PS II and PS I [40].
The FV/Fm values of the two strains showed no significant difference as the temperature increased from 20 °C to 30 °C. However, heat stress resulted in an apparent decrease in the PIabs value in both wild type and mutant. These results suggested that PIabs was more sensitive to stress compared to FV/Fm, which was consistent with previous reports [41]. PIabs is a multiparametric expression and was highly used to identify plant vitality exposed to environmental stress, and it takes into consideration the three main functional steps: the density of active reaction centers per chlorophyll; the ratio of the de-excitation rate constants for photochemical and non-photochemical events; and the efficiency of the conversion of excitation energy to electron transport toward the plastoquinone pool [21,42]. A large decrease in PIabs of the wild type indicated the destructive effects of heat stress on PSII. While the small decrease in the PIabs value in the green mutant indicated that the PSII overall performance was less affected by heat stress in the green mutant than in the wild type.
Heat stress can block PSII reaction centers (RCs) and dissociate antenna pigment–protein complexes from the PSII light-harvesting complex [43,44,45]. Among the PSII partial complexes, OEC is more sensitive to heat [46]. In the present study, OEC inactivation was observed in the wild type after exposure to heat stress, as evidenced by the increase in WK in the wild type. RC/CSo decreased in the wild type under heat stress, which meant that the amount of active PSII RCs greatly decreased. In addition, the decrease in φEo and φRo was more in the wild type than in the green mutant under heat stress, indicating higher inhibition of PSII electron transport. In contrast, in the green mutant, OEC was less affected by high temperature, and the lower decrease in active PSII RCs and lower inhibition of electron transport in response to high temperature could help the green mutant adapt to heat stress.
Heat dissipation is essential for protecting the leaf from stress-induced damage [28]. The NPQ process can harmlessly dissipate the excess energy, thus alleviating photodamage by reducing ROS production [47,48,49]. In the present study, the lower increase in NPQ in the wild type indicated lower heat dissipation capacity in response to heat stress, which may cause ROS accumulation. In contrast, the green mutant could be able to dissipate heat efficiently, reducing the risk of oxidative stress attacking the photosynthetic apparatus. ROS production, including single oxygen (1O2), superoxide (O2), hydrogen peroxide (H2O2), and hydroxyl radical (OH) production, is a sign of cellular injury caused by heat stress [3]. Superoxide, the product of a one-electron reduction in oxygen, is a ROS precursor and a mediator in oxidative chain reactions [50]. The superoxide content, measured as OFR, can reflect the ROS level in organisms. In this study, the OFR of the wild type increased much more than that of the green mutant under heat stress, indicating that it was subjected to severe oxidative stress. ROS can cause the autocatalytic peroxidation of membrane lipids, which is indicated by MDA, a product of the reaction [3]. After exposure to heat stress, MDA accumulated to a greater extent in the wild type than in the green mutant. The OFR and MDA of the green mutant showed a much lower increase under heat stress than the wild type, indicating its superior ability to maintain cell membrane integrity and minimize oxidative damage. To scavenge ROS, cellular antioxidant defenses that produce more antioxidant enzymes, such as SOD, POD, and CAT, are usually activated [6]. In this study, POD and CAT activities in the two strains increased after heat stress. The high POD and CAT activities in the green mutant suggest that it possessed an efficient antioxidant system for protection from oxidative stress and was able to inhibit ROS production in response to heat stress. The enhanced ROS detoxification capacity allowed the green mutant to better maintain its photosynthesis ability.

3.2. Low Temperature Stress in G. lemaneiformis

Chilling temperature can affect algal photosynthesis [51]. Low temperature inhibits carbon dioxide fixation, leading to ROS production, which in turn suppresses de novo D1 protein synthesis and reduces PSII repair, resulting in more severe damage to PSII [52]. In the present study, the OJIP fluorescence intensity of the two strains was lower under low temperature, suggesting the inhibition of photosynthetic performance. The greater decrease in FV/Fm and PIabs values in the green mutant indicated that it had a greater degree of inhibition of photosynthetic performance.
However, the increase in OFR content was greater in the wild type than in the green mutant under low-temperature conditions, indicating that the green mutant has lower ROS levels than the wild type. Excessive ROS accumulation induces the production of aldehydes, which can cause genotoxic effects, such as lipid peroxidation [53]. The level of lipid peroxidation was much higher in the wild type than in the green mutant. The ROS and peroxidation levels of the green mutant remained low and increased after a low-temperature treatment, suggesting that the antioxidant potential of the green mutant may be higher under low-temperature conditions. The trend of increasing SOD activity was higher in the green mutant than in the wild type under low-temperature conditions. SOD is the first line of defense against ROS; the enzyme converts superoxide to hydrogen peroxide and oxygen [54]. The superoxide content and lipid peroxidation level were also lower in the green mutant than in the wild type under low temperatures. An increase in SOD activity was observed in G. lemaneiformis grown under low temperature, suggesting that reduced temperature enhances antioxidant defense [6]. There may be some mechanism in the green mutant to cope with the low-temperature stress, which deserves further investigation.

4. Materials and Methods

4.1. Sample Collection, Treatment, and Identification

The two G. lemaneiformis strains were obtained from cultivation rafts in Gaolv Aquaculture Co., Ltd., Rongcheng, Shandong, China (122.6° E, 37.2° N). The stable green-pigmented mutant was derived from the G. lemaneiformis cultivar 981 by spontaneous mutation and continuous selection for 3 years from 2019. After collection, the two G. lemaneiformis strains were first cultured in conical flasks with filtered seawater at 20 °C for 5 days at an irradiance of 72 μmol photons m−2 s−1 (12 h light:12 h dark).
ITS sequences were obtained for the green mutant and the wild type (through DNA extraction, PCR amplification, and sequencing) in addition to the seven sequences from GenBank. The evolutionary history was inferred by using the maximum likelihood method based on the general time reversible model [55]. The tree with the highest log likelihood (−6831.85) is shown. The percentage of trees in which the associated taxa clustered together is shown next to the branches. Initial tree(s) for the heuristic search were obtained automatically by applying Neighbor-Join and BioNJ algorithms to a matrix of pairwise distances estimated using the Maximum Composite Likelihood (MCL) approach and then selecting the topology with superior log likelihood value. A discrete Gamma distribution was used to model evolutionary rate differences among sites (five categories (+G, parameter = 1.9160)). The rate variation model allowed for some sites to be evolutionarily invariable ([+I], 16.84% sites). The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. Codon positions included were 1st+2nd+3rd+Noncoding. There were a total of 1171 positions in the final dataset. Evolutionary analyses were conducted in MEGA7 [56].

4.2. Experimental Design

For temperature treatments, the two G. lemaneiformis strains were cultured at three temperatures—low (8 °C), medium (20 °C), and high (30 °C). The experimental temperature was selected according to the actual culture temperature of G. lemaneiformis. The culture was started when 3.0 g of fresh-weight algae was introduced into 18 conical flasks containing 1 L filtered seawater. Light conditions (intensity and period) for all treatments were the same as indicated above. Samples were cultivated for 7 days and harvested to assess physiological and biochemical responses under different temperature conditions. Replicate cultures (n = 3) were maintained at each treatment condition to avoid pseudoreplication.

4.3. Chlorophyll Fluorescence Measurements

Fast chlorophyll fluorescence intensity of the OJIP transient was measured under the FC1000-H fluorescence imaging system (Photon Systems Instruments Ltd., Brno, Czech Republic). All G. lemaneiformis samples were kept in the dark for 15–20 min before measurement to keep the PSII reaction centers fully open. The fluorescence intensity of the O, J, I, and P steps was recorded after 20 μs (FO), 2 ms (FJ), 30 ms (FI), and 1000 ms (FP), respectively. The saturated light intensity of the OJIP transient was 3000 μmol m−2 s−1. Abbreviations, formulae, and definitions of the JIP-test parameters are listed in Table 1 [57,58,59].

4.4. Enzyme Activity Measurements

Oxygen free radical (OFR) and malondialdehyde (MDA) content, and SOD, CAT, and POD activities were measured using physiological assay kits (Suzhou Grace Biotechnology Co., Ltd., Suzhou, China) according to the manufacturers’ recommendations. All processes were biologically and temporally repeated in three independent and parallel experiments.

4.5. Statistics

All data are expressed as means ± standard deviation (SD, n = 3). Statistical analyses were performed using SPSS v.21.0 (IBM Corp., Armonk, NY, USA).The differences between treatments (HT and LT) and the control (MT) were considered significant if p < 0.05.

5. Conclusions

Compared with the wild type of G. lemaneiformis, the green mutant showed more tolerance to heat stress, as indicated by its higher photosynthetic performance and stronger antioxidant activity. In addition, the green mutant produced less ROS under low temperatures, which may have some potential to alleviate the damage caused by a decrease in photosynthetic performance. In conclusion, the green mutant has increased tolerance to high temperatures, laying the foundation for the selection of a resistant strain of G. lemaneiformis, while the green mutant may provide material for further genetic research and production applications as a superior algal strain.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants12081670/s1, Table S1: JIP-test parameters deduced from chlorophyll a fluorescence OJIP transient curves in the green mutant and wild type of G. lemaneiformis exposed to three temperature levels for 7 days.

Author Contributions

X.L., X.M. and X.Y., were responsible for experiments, data analysis, interpretation, and manuscript preparation; D.D. gave a contribution to the experimental design and editing of the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Asia Collaboration Project on Development of Ecological Marine Ranching.

Data Availability Statement

The datasets used and/or analyzed during this study are available from the corresponding author upon reasonable request.

Acknowledgments

We thank the support of the Asia Collaboration Project on Development of Ecological Marine Ranching for this manuscript.

Conflicts of Interest

All the authors have approved the manuscript and agree with submission to your esteemed journal. There are no conflicts of interest to declare.

References

  1. Mantri, V.A.; Kambey, C.S.B.; Cottier-Cook, E.J.; Usandizaga, S.; Buschmann, A.H.; Chung, I.K.; Liu, T.; Sondak, C.F.A.; Qi, Z.; Lim, P.E.; et al. Overview of global Gracilaria production, the role of biosecurity policies and regulations in the sustainable development of this industry. Rev. Aquac. 2022, 15, 801–819. [Google Scholar] [CrossRef]
  2. Bureau of Fisheries, Ministry of Agriculture and Rural Affairs. China Fishery Statistics Yearbook; China Agriculture Press: Beijing, China, 2021; pp. 23–28.
  3. Fu, F.; Sui, Z.H.; Zhou, W.; Wang, J.G.; Chang, L.P.; Ci, S.F. UV-irradiation mutation of tetraspores of Gracilariopsis lemaneiformis and screening of thermotolerant strains. J. Appl. Phycol. 2013, 26, 647–656. [Google Scholar] [CrossRef]
  4. Fei, X.G. Solving the coastal eutrophication problem by large scale seaweed cultivation. Hydrobiologia 2004, 512, 145–151. [Google Scholar] [CrossRef]
  5. Zhou, Y.; Yang, H.; Hu, H.; Liu, Y.; Mao, Y.; Zhou, H.; Xu, X.; Zhang, F. Bioremediation potential of the macroalga Gracilaria lemaneiformis (Rhodophyta) integrated into fed fish culture in coastal waters of north China. Aquaculture 2006, 252, 264–276. [Google Scholar] [CrossRef]
  6. Liu, L.; Zou, D.; Jiang, H.; Chen, B.; Zeng, X. Effects of increased CO2 and temperature on the growth and photosynthesis in the marine macroalga Gracilaria lemaneiformis from the coastal waters of South China. J. Appl. Phycol. 2017, 30, 1271–1280. [Google Scholar] [CrossRef]
  7. Shi, C.S.; Sang, Y.X.; Sun, G.Q.; Li, T.Y.; Gong, Z.S.; Wang, X.H. Characterization and bioactivities of a novel polysaccharide obtained from Gracilariopsis lemaneiformis. An. Acad. Bras. Cienc. 2017, 89, 175–189. [Google Scholar] [CrossRef]
  8. Wang, X.; Zhang, Z.; Wu, Y.; Sun, X.; Xu, N. Synthesized sulfated and acetylated derivatives of polysaccharide extracted from Gracilariopsis lemaneiformis and their potential antioxidant and immunological activity. Int. J. Biol. Macromol. 2019, 124, 568–572. [Google Scholar] [CrossRef]
  9. Fang, T.; Zhang, X.; Hu, S.; Yu, Y.; Sun, X.; Xu, N. Enzymatic degradation of Gracilariopsis lemaneiformis polysaccharide and the antioxidant activity of its degradation products. Mar. Drugs 2021, 19, 270. [Google Scholar] [CrossRef]
  10. Nauer, F.; Oliveira, M.C.; Plastino, E.M.; Yokoya, N.S.; Fujii, M.T. Coping with heatwaves: How a key species of seaweed responds to heat stress along its latitudinal gradient. Mar. Environ. Res. 2022, 177, 105620. [Google Scholar] [CrossRef]
  11. Zhang, X.; Hu, C.; Sun, X.; Zang, X.; Zhang, X.; Fang, T.; Xu, N. Comparative transcriptome analysis reveals chitooligosaccharides-induced stress tolerance of Gracilariopsis lemaneiformis under high temperature stress. Aquaculture 2020, 519, 734876. [Google Scholar] [CrossRef]
  12. Zhang, X.C.; Fei, X.G.; Wang, G.C.; Lin, X.Z.; Chen, W.Z.; Sui, Z.H.; Xu, D.; Zang, X.N. Genetic studies and large scale cultivation of Gracilaria lemaneiformis. J. Ocean Univ. China 2009, 39, 947–954. [Google Scholar]
  13. Gu, Y.; Zhang, X.; Lu, N.; Zang, X.; Zhang, X.; Li, G. Cloning and transcription analysis of hsp70-1 and hsp70-2 of Gracilaria lemaneiformis under heat shock. Aquaculture 2012, 358–359, 284–291. [Google Scholar] [CrossRef]
  14. Egan, S.; Fernandes, N.D.; Kumar, V.; Gardiner, M.; Thomas, T. Bacterial pathogens, virulence mechanism and host defence in marine macroalgae. Environ. Microbiol. 2014, 16, 925–938. [Google Scholar] [CrossRef]
  15. Wang, Y.; Feng, Y.; Wang, H.; Zhong, M.; Chen, W.; Du, H. Physiological and proteomic analyses of two Gracilaria lemaneiformis strains in response to high-temperature stress. J. Appl. Phycol. 2015, 28, 1847–1858. [Google Scholar] [CrossRef]
  16. Qin, F.; Zang, X.; Shui, G.; Wang, Z. Transcriptome analysis of Gracilariopsis lemaneiformis at low temperature. J. Appl. Phycol. 2021, 33, 4035–4050. [Google Scholar] [CrossRef]
  17. Yang, Y.F.; Fei, X.G.; Song, J.M.; Hu, H.Y.; Wang, G.C.; Chung, I.K. Growth of Gracilaria lemaneiformis under different cultivation conditions and its effects on nutrient removal in Chinese coastal waters. Aquaculture 2006, 254, 248–255. [Google Scholar] [CrossRef]
  18. Larosa, V.; Meneghesso, A.; La Rocca, N.; Steinbeck, J.; Hippler, M.; Szabo, I.; Morosinotto, T. Mitochondria affect photosynthetic electron transport and photosensitivity in a green alga. Plant Physiol. 2018, 176, 2305–2314. [Google Scholar] [CrossRef]
  19. Huot, Y.; Babin, M. Overview of fluorescence protocols: Theory, basic concepts, and Practice. In Chlorophyll a Fluorescence in Aquatic Sciences: Methods and Applications; Suggett, D., Prášil, O., Borowitzka, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 4, pp. 31–75. [Google Scholar] [CrossRef]
  20. Danielsson, R.; Suorsa, M.; Paakkarinen, V.; Albertsson, P.A.; Styring, S.; Aro, E.M.; Mamedov, F. Dimeric and monomeric organization of photosystem II. Distribution of five distinct complexes in the different domains of the thylakoid membrane. J. Biol. Chem. 2006, 281, 14241–14249. [Google Scholar] [CrossRef]
  21. Strasser, R.J.; Tsimilli-Michael, M.; Srivastava, A. Analysis of the chlorophyll a fluorescence transient. In Chlorophyll Fluorescence a Signature of Photosynthesis; Advances in Photosynthesis and Respiration; Papageorgiou, G.C., Govindjee, X., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2004; Volume 19, pp. 321–362. [Google Scholar]
  22. Stirbet, A.; Govindjee. On the relation between the Kautsky effect (chlorophyll a fluorescence induction) and Photosystem II: Basics and applications of the OJIP fluorescence transient. J. Photochem. Photobiol. B 2011, 104, 236–257. [Google Scholar] [CrossRef]
  23. Maxwell, K.; Johnson, G.N. Chlorophyll fluorescence—A practical guide. J. Exp. Bot. 2000, 51, 659–668. [Google Scholar] [CrossRef]
  24. Baker, N.R. Chlorophyll fluorescence: A probe of photosynthesis in vivo. Annu. Rev. Plant. Biol. 2008, 59, 89–113. [Google Scholar] [CrossRef] [PubMed]
  25. Guo, H.; Yao, J.; Sun, Z.; Duan, D. Effects of salinity and nutrients on the growth and chlorophyll fluorescence of Caulerpa lentillifera. J. Oceanol. Limnol. 2014, 33, 410–418. [Google Scholar] [CrossRef]
  26. Kang, J.W.; Kambey, C.; Shen, Z.; Yang, Y.F.; Chung, I.K. The short-term effects of elevated CO2 and ammonium concentrations on physiological responses in Gracilariopsis lemaneiformis (Rhodophyta). Fish. Aquat. Sci. 2017, 20, 18. [Google Scholar] [CrossRef]
  27. Ye, C.P.; Zhang, M.C.; Zhao, J.G.; Yang, Y.F.; Zuo, Y. Photosynthetic response of the macroalga, Gracilaria lemaneiformis(Rhodophyta), to various N and P levels at different temperatures. Int. Rev. Hydrobiol. 2013, 98, 245–252. [Google Scholar] [CrossRef]
  28. Wu, H.; Jiang, H.; Liu, C.; Deng, Y. Growth, pigment composition, chlorophyll fluorescence and antioxidant defenses in the red alga Gracilaria lemaneiformis (Gracilariales, Rhodophyta) under light stress. South Afr. J. Bot. 2015, 100, 27–32. [Google Scholar] [CrossRef]
  29. Müller, R.; Desel, C.; Steinhoff, F.S.; Wiencke, C.; Bischof, K. UV-radiation and elevated temperatures induce formation of reactive oxygen species in gametophytes of cold-temperate/Arctic kelps (Laminariales, Phaeophyceae). Phycol. Res. 2012, 60, 27–36. [Google Scholar] [CrossRef]
  30. Lu, Y.; Yao, J. Chloroplasts at the crossroad of photosynthesis, pathogen infection and plant defense. Int. J. Mol. Sci. 2018, 19, 3900. [Google Scholar] [CrossRef]
  31. Ledford, H.K.; Niyogi, K.K. Singlet oxygen and photo-oxidative stress management in plants and algae. Plant Cell Environ. 2005, 28, 1037–1045. [Google Scholar] [CrossRef]
  32. Kim, C. ROS-driven oxidative modification: Its impact on chloroplasts-nucleus communication. Front. Plant Sci. 2019, 10, 1729. [Google Scholar] [CrossRef]
  33. Zhao, D.K.; Wang, H.P.; Chen, S.Y.; Yu, D.Q.; Reiter, R.J. Phytomelatonin: An emerging regulator of plant biotic stress resistance. Trends Plant Sci. 2021, 26, 70–82. [Google Scholar] [CrossRef]
  34. Fu, F.; Sui, Z.H.; Sun, L.Q.; Bi, L.P.; Ding, L.J.; Chen, Q. Research advance on the algal mutation breeding technologies. Biotechnol. Bull. 2018, 34, 58–63. [Google Scholar] [CrossRef]
  35. Chen, W.Z.; Xu, D.; Wang, L.G.; Meng, L.; Du, H.; Zhang, X.C. Preliminary study on economic characteristics and agar characteristics of two new strains of Gracilaria lemaneiformis. J. Ocean Univ. China 2009, 39, 437–442. [Google Scholar] [CrossRef]
  36. Jiang, M.J.; Bi, G.Q.; Wang, J.G.; Zhang, J.Y.; Hu, Y.Y.; Li, X.D.; Liu, J.; Sui, Z.H. WGCN: Aanalysis of the development of the tetrasporohyte of Graci lariopsislemaneiformis (Rhodophyta). J. Ocean Univ. China 2020, 50, 61–75. [Google Scholar] [CrossRef]
  37. Huang, X.Y.; Zang, X.N.; Wu, F.; Jin, Y.M.; Wang, H.T.; Liu, C.; Ding, Y.T.; He, B.X.; Xiao, D.F.; Song, X.W.; et al. Transcriptome sequencing of Gracilariopsis lemaneiformis to analyze the genes related to optically active phycoerythrin synthesis. PLoS ONE 2017, 12, e0170855. [Google Scholar] [CrossRef] [PubMed]
  38. Lin, L.C.; Lu, Y.; Sun, P.; Sun, X.; Xu, N.J. Response of two chotismate metabolic enzymes to temperature and salicylic acid in Gracilariopsis lemaneiformis and their prokaryotic expression analysis. Oceanol. Limnol. Sin. 2019, 50, 220–227. [Google Scholar] [CrossRef]
  39. Jarvi, S.; Suorsa, M.; Aro, E.M. Photosystem II repair in plant chloroplasts--regulation, assisting proteins and shared components with photosystem II biogenesis. Biochim. Biophys. Acta 2015, 1847, 900–909. [Google Scholar] [CrossRef] [PubMed]
  40. Strasser, R.; Srivastava, A.; Tsimilli-Michael, M. The fluorescence transient as a tool to characterize and screen photosynthetic samples. In Probing Photosynthesis: Mechanisms, Regulation and Adaptation; Yunus, M., Pathre, U., Mohanty, P., Eds.; Taylor & Francis Publishers: London, UK, 2000; pp. 445–483. [Google Scholar]
  41. Zushi, K.; Kajiwara, S.; Matsuzoe, N. Chlorophyll a fluorescence OJIP transient as a tool to characterize and evaluate response to heat and chilling stress in tomato leaf and fruit. Sci. Hortic. 2012, 148, 39–46. [Google Scholar] [CrossRef]
  42. Oukarroum, A.; Madidi, E.S.; Strasser, R.J. Differential heat sensitivity index in barley cultivars (Hordeum vulgare L.) monitored by chlorophyll a fluorescence OKJIP. Plant Physiol. Biochem. 2016, 105, 102–108. [Google Scholar] [CrossRef]
  43. Takahashi, S.; Murata, N. How do environmental stresses accelerate photoinhibition? Trends Plant Sci. 2008, 13, 178–182. [Google Scholar] [CrossRef]
  44. Takahashi, S.; Milward, S.E.; Fan, D.Y.; Chow, W.S.; Badger, M.R. How does cyclic electron flow alleviate photoinhibition in Arabidopsis? Plant Physiol. 2009, 149, 1560–1567. [Google Scholar] [CrossRef]
  45. Sarvikas, P.; Tyystjarvi, T.; Tyystjarvi, E. Kinetics of prolonged photoinhibition revisited: Photoinhibited Photosystem II centres do not protect the active ones against loss of oxygen evolution. Photosynth. Res. 2010, 103, 7–17. [Google Scholar] [CrossRef] [PubMed]
  46. Heckathorn, S.A.; Coleman, J.S.; Hallberg, R.L. Recovery of net CO2 assimilation after heat stress is correlated with recovery of oxygen-evolving-complex proteins in Zea mays L. Photosynthetica 1997, 34, 13–20. [Google Scholar] [CrossRef]
  47. Hideg, E.; Kos, P.B.; Schreiber, U. Imaging of NPQ and ROS formation in tobacco leaves: Heat inactivation of the water-water cycle prevents down-regulation of PSII. Plant Cell Physiol. 2008, 49, 1879–1986. [Google Scholar] [CrossRef] [PubMed]
  48. Zhao, X.; Chen, T.; Feng, B.; Zhang, C.; Peng, S.; Zhang, X.; Fu, G.; Tao, L. Non-photochemical quenching plays a key role in light acclimation of rice plants differing in leaf color. Front. Plant Sci. 2016, 7, 1968. [Google Scholar] [CrossRef]
  49. Moustakas, M.; Moustaka, J.; Sperdouli, I. Hormesis in photosystem II: A mechanistic understanding. Curr. Opin. Toxicol. 2022, 29, 57–64. [Google Scholar] [CrossRef]
  50. Turrens, J.F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003, 552, 335–344. [Google Scholar] [CrossRef]
  51. Terada, R.; Matsumoto, K.; Borlongan, I.A.; Watanabe, Y.; Nishihara, G.N.; Endo, H.; Shimada, S. The combined effects of PAR and temperature including the chilling-light stress on the photosynthesis of a temperate brown alga, Sargassum patens (Fucales), based on field and laboratory measurements. J. Appl. Phycol. 2017, 30, 1893–1904. [Google Scholar] [CrossRef]
  52. Allakhverdiev, S.I.; Murata, N. Environmental stress inhibits the synthesis de novo of proteins involved in the photodamage-repair cycle of Photosystem II in Synechocystis sp. PCC 6803. Biochim. Biophys. Acta 2004, 1657, 23–32. [Google Scholar] [CrossRef]
  53. Uji, T.; Mizuta, H. 1-Aminocyclopropane-1-carboxylic acid and its analogs alleviate heat stress damage in the marine red alga Neopyropia yezoensis (Rhodophyta). J. Appl. Phycol. 2022, 34, 1527–1536. [Google Scholar] [CrossRef]
  54. Machado, M.D.; Soares, E.V. Exposure of the alga Pseudokirchneriella subcapitata to environmentally relevant concentrations of the herbicide metolachlor: Impact on the redox homeostasis. Ecotoxicol. Environ. Saf. 2021, 207, 111264. [Google Scholar] [CrossRef]
  55. Nei, M.; Kumar, S. Molecular Evolution and Phylogenetics; Oxford University Press: New York, NY, USA, 2000. [Google Scholar]
  56. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, S.; Yang, J.; Zhang, M.; Strasser, R.J.; Qiang, S. Classification and characteristics of heat tolerance in Ageratina adenophora populations using fast chlorophyll a fluorescence rise O-J-I-P. Environ. Exp. Bot. 2016, 122, 126–140. [Google Scholar] [CrossRef]
  58. Yang, X.Q.; Zhang, Q.S.; Zhang, D.; Sheng, Z.T. Light intensity dependent photosynthetic electron transport in eelgrass (Zostera marina L.). Plant Physiol. Biochem. 2017, 113, 168–176. [Google Scholar] [CrossRef]
  59. Zeng, F.L.; Wang, G.J.; Liang, Y.P.; Guo, N.H.; Zhu, L.; Wang, Q.; Chen, H.W.; Ma, D.R.; Wang, J.Y. Disentangling the photosynthesis performance in japonica rice during natural leaf senescence using OJIP fluorescence transient analysis. Funct. Plant Biol. 2021, 48, 206–217. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Two G. lemaneiformis strains. Green-pigmented mutant (a) and wild type (b).
Figure 1. Two G. lemaneiformis strains. Green-pigmented mutant (a) and wild type (b).
Plants 12 01670 g001
Figure 2. Molecular phylogenetic analysis by maximum likelihood method based on internal transcribed spacer (ITS).
Figure 2. Molecular phylogenetic analysis by maximum likelihood method based on internal transcribed spacer (ITS).
Plants 12 01670 g002
Figure 3. OJIP transient curves in the green mutant (a) and wild type (b) of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C (blue). MT, medium temperature 20 °C (black). HT, high temperature 30 °C (red). MT was the control.
Figure 3. OJIP transient curves in the green mutant (a) and wild type (b) of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C (blue). MT, medium temperature 20 °C (black). HT, high temperature 30 °C (red). MT was the control.
Plants 12 01670 g003
Figure 4. Spider plots of JIP-test parameters deduced from chlorophyll a fluorescence OJIP transient curves in the green mutant (a) and wild type (b) of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C (blue). MT, medium temperature 20 °C (black). HT, high temperature 30 °C (red). The differences between treatments (HT and LT) and the control (MT) were considered significant at * p < 0.05.
Figure 4. Spider plots of JIP-test parameters deduced from chlorophyll a fluorescence OJIP transient curves in the green mutant (a) and wild type (b) of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C (blue). MT, medium temperature 20 °C (black). HT, high temperature 30 °C (red). The differences between treatments (HT and LT) and the control (MT) were considered significant at * p < 0.05.
Plants 12 01670 g004
Figure 5. Non-photochemical quenching in the green mutant and wild type of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C. MT, medium temperature 20 °C. HT, high temperature 30 °C. The differences between treatments (HT and LT) and the control (MT) were considered significant at ** p < 0.01.
Figure 5. Non-photochemical quenching in the green mutant and wild type of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C. MT, medium temperature 20 °C. HT, high temperature 30 °C. The differences between treatments (HT and LT) and the control (MT) were considered significant at ** p < 0.01.
Plants 12 01670 g005
Figure 6. Superoxide content (OFR, (a)) and lipid peroxidation level (MDA, (b)) of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C. MT, medium temperature 20 °C (control). HT, high temperature 30 °C. The differences between treatments (HT and LT) and the control (MT) were considered significant at ** p < 0.01.
Figure 6. Superoxide content (OFR, (a)) and lipid peroxidation level (MDA, (b)) of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C. MT, medium temperature 20 °C (control). HT, high temperature 30 °C. The differences between treatments (HT and LT) and the control (MT) were considered significant at ** p < 0.01.
Plants 12 01670 g006
Figure 7. Superoxide dismutase (SOD, (a)), peroxidase (POD, (b)), and catalase (CAT, (c)) activities in the two strains of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C. MT, medium temperature 20 °C (control). HT, high temperature 30 °C. The differences between treatments (HT and LT) and the control (MT) were considered significant at ** p < 0.01.
Figure 7. Superoxide dismutase (SOD, (a)), peroxidase (POD, (b)), and catalase (CAT, (c)) activities in the two strains of G. lemaneiformis exposed to three temperature levels for 7 days. LT, low temperature 8 °C. MT, medium temperature 20 °C (control). HT, high temperature 30 °C. The differences between treatments (HT and LT) and the control (MT) were considered significant at ** p < 0.01.
Plants 12 01670 g007
Table 1. Abbreviations, formulae, and definitions of JIP-test parameters.
Table 1. Abbreviations, formulae, and definitions of JIP-test parameters.
ParametersDefinitions
FO = F20μsMinimum fluorescence, all PSII RCs are open
FK = F300μsFluorescence intensity at the K-step (300 µs) of OJIP
FJ = F2msFluorescence intensity at the J-step (2 ms) of OJIP
FI = F30msFluorescence intensity at the I-step (30 ms) of OJIP
Fm = FPMaximum fluorescence at peak P of OJIP
VJ = (FJFO)/(FmFO)Relative variable fluorescence at the J-step of the fluorescence induction curve
VI = (FIFO)/(FmFO)Relative variable fluorescence at the I-step of the fluorescence induction curve
VK = (FKFO)/(FmFO)Relative variable fluorescence at the K-step of the fluorescence induction curve
WK = (FKFO)/(FJFO)Damage to the oxygen-evolving complex
Mo = 4(FKFO)/(FmFO)Approximated initial slope (in ms−1) of the fluorescence transient
FV/FmMaximum photochemical quantum yield
φPo = 1 − FO/Fm = FV/FmMaximum quantum yield for primary photochemistry
φEo = FV/Fm × (1− VJ)Quantum yield for electron transport
φRo = FV/Fm × (1− VI)Quantum yield of reduction of end electron acceptors of PSI
Ψo = 1 − VJEfficiency with which a trapped exaction can move an electron into the electron transport chain further than QA
ABS/RC = (Mo/VJ)/(1/φPo)Light absorption flux (for PSII antenna chlorophyll) per RC
ETo/RC = (Mo/VJ) × (1 − VJ)Maximum electron transport flux (further than QA) per PSII RC
TRo/RC = (Mo/VJ)Trapped (maximum) energy flux (leading to QA reduction) per RC
DIo/RC = ABS/RC − TRo/RCDissipation energy flux per PSII RC
RC/CSo = φPo × (VJ/Mo) × (ABS/CS)Number of active RCs per CS
PIabs = (RC/ABS) × [φPo/(1 − φPo)] × [ψo/(1 − ψo)]Performance index on absorption basis
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, X.; Meng, X.; Yang, X.; Duan, D. Characterization of Chlorophyll Fluorescence and Antioxidant Defense Parameters of Two Gracilariopsis lemaneiformis Strains under Different Temperatures. Plants 2023, 12, 1670. https://doi.org/10.3390/plants12081670

AMA Style

Li X, Meng X, Yang X, Duan D. Characterization of Chlorophyll Fluorescence and Antioxidant Defense Parameters of Two Gracilariopsis lemaneiformis Strains under Different Temperatures. Plants. 2023; 12(8):1670. https://doi.org/10.3390/plants12081670

Chicago/Turabian Style

Li, Xiaomei, Xue Meng, Xiaoqi Yang, and Delin Duan. 2023. "Characterization of Chlorophyll Fluorescence and Antioxidant Defense Parameters of Two Gracilariopsis lemaneiformis Strains under Different Temperatures" Plants 12, no. 8: 1670. https://doi.org/10.3390/plants12081670

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop