Next Article in Journal
Integrative Meta-Analysis of Huntington’s Disease Transcriptome Landscape
Next Article in Special Issue
Stage-Specific Transcriptomes of the Mussel Mytilus coruscus Reveals the Developmental Program for the Planktonic to Benthic Transition
Previous Article in Journal
Genetic Diversity and Population Structure Analysis of Castanopsis hystrix and Construction of a Core Collection Using Phenotypic Traits and Molecular Markers
Previous Article in Special Issue
Comparing the Efficiency of Single-Locus Species Delimitation Methods within Trochoidea (Gastropoda: Vetigastropoda)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

AMPK Promotes Larval Metamorphosis of Mytilus coruscus

1
Shanghai Collaborative Innovation Center for Cultivating Elite Breeds and Green-Culture of Aquaculture animals, Shanghai 201306, China
2
International Research Center for Marine Biosciences, Ministry of Science and Technology, Shanghai Ocean University, Shanghai 201306, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2022, 13(12), 2384; https://doi.org/10.3390/genes13122384
Submission received: 10 November 2022 / Revised: 30 November 2022 / Accepted: 9 December 2022 / Published: 16 December 2022
(This article belongs to the Special Issue Genetic Breeding and Genomics of Marine Shellfish)

Abstract

:
Metamorphosis is a critical process in the transition from planktonic life to benthic life for marine invertebrates, which is accompanied by a large amount of energy consumption. Previous studies have proved that AMP-activated protein kinase (AMPK), as a vital energy regulator, plays a prominent role in mediating the growth and development of terrestrial animals. However, its function in the growth and development of marine invertebrates, especially in metamorphosis, remains elusive. This study explored the function of AMPK in the larval metamorphosis of Mytilus coruscus. The full-length cDNA of AMPK genes in M. coruscus was cloned and characterized, which is composed of three subunits, McAMPKα, McAMPKβ, and McAMPKγ. Pharmacological tests demonstrated that through the application of an AMPK activator, AMP substantially enhanced the larval metamorphosis rate (p < 0.05). By contrast, the larval metamorphosis rate decreased significantly after being treated with the AMPK inhibitor Compound C (p < 0.05). McAMPK gene knock-down resulted in a reduction in McAMPK gene expression (p < 0.05), and the larval metamorphosis of M. coruscus was significantly restrained (p < 0.05). These results indicated that AMPK signaling is vital in the larval metamorphosis of M. coruscus, which advances further understanding in exploring the molecular mechanisms in the metamorphosis of marine invertebrate larvae.

1. Introduction

Mytilus coruscus is extensively spread along the coastline of East Asia and is one of the main economical shellfish for aquaculture in China [1]. Nevertheless, the wild populations of M. coruscus have decreased tremendously due to human over-harvesting and marine environmental changes in recent years. Thus, most of the seedlings are raised indoors, and it is still difficult to break the technical bottleneck of the high mortality of artificial seedlings during larval metamorphosis, which restricts the sustainable development of the hard-shelled mussel breeding industry [2]. Like most marine invertebrates, M. coruscus undergoes a crucial process of metamorphosis to complete the transition from the planktonic stage to the benthic stage for survival. Thus, resolving the metamorphosis mechanism of M. coruscus larvae is conducive to the mussel aquaculture industry.
Metamorphosis means a special transformation in the life cycle of animals from larvae to juveniles, during which the morphology, physiology, and ecology of creatures will be adjusted profoundly [3]. This meaningful event occurs throughout the animal kingdom; amphibians [4], fish [5], and terrestrial and aquatic invertebrates [6,7,8] all undergo metamorphosis to survive better during their life. During the metamorphosis process, the larvae go through many drastic morphological changes; for some marine mussels, this includes the disappearance of the face plate, the shedding of the ciliary ring, the rapid growth of gill filaments, and the production of adult shells so as to develop to post-larvae and grow into adults, which need to consume a large amount of energy [9]. Some investigations have indicated that α2-adrenergic receptors [6], nitric oxide [10], adenosine kinase [11], and other signals can govern the metamorphosis process of marine mussel larvae; however, the mechanism of metamorphosis remains poorly understood.
As a master molecular sensor, AMP-activated protein kinase (AMPK) is fundamental in regulating cell energy homeostasis, organism development, and tissue growth [12,13,14]. Eukaryotes possess a highly elegant mechanism for detecting insufficient cellular ATP levels through the serine/threonine kinase AMPK compound [14], which is made up of one α catalytic subunit (α1 and α2) and two regulatory subunits, β (β1 and β2) and γ (γ1, γ2, and γ3), respectively [14,15]. When cells encounter energy stress, AMPK serves as a metabolic sensor, stimulating catabolic processes to produce additional ATP while suppressing anabolism to achieve energy balance [14,16,17]. In other words, AMPK maintains the balance of energy metabolism by regulating the synthesis and transformation of fat, protein, and glucose [18,19,20,21]. Furthermore, AMPK is in charge of maintaining germline quiescence in Caenorhabditis elegans larvae by changing the germline chromatin landscape to preserve germ cell integrity and regulating larvae metabolism to promote their smooth growth into adults [22]. When AMPK is activated, it enhances oxidative phosphorylation and the usage of pyruvate and allows Drosophila melanogaster to prune their dendrites of sensory neurons to complete developmental transitions [23].
Metamorphosis is a vital process in the life of marine invertebrates, which is accompanied by huge energy consumption. Research has shown that protein contributes the most energy to oyster larval metamorphosis, accounting for 69.3% of the total energy, followed by fat (24.3%) and sugar (6.4%) [24]. Moreover, proteins involved in glycolysis and fatty acid metabolism were upregulated during the metamorphosis of Balanus amphitrite [25]. This evidence indicates that energy metabolism plays a significant role in larval metamorphosis. As a fundamental energy-regulating signal, the role of AMPK in insect metamorphosis has been revealed, but the function of AMPK in the metamorphosis of marine invertebrates is rarely reported, especially in M. coruscus.
In this study, the full-length cDNA of AMPK in M. coruscus was cloned and characterized, which is composed of three subunits, named McAMPKα, McAMPKβ, and McAMPKγ. Pharmacology and RNA interference (RNAi) experiments demonstrated that when AMPK signaling was inhibited, the larval metamorphosis was restrained, whereas activating AMPK signaling promoted larval metamorphosis. This indicates that AMPK is essential for the metamorphosis of M. coruscus larvae. Given the significance of M. coruscus in the commercial bivalve breeding industry of China, this study casts the spotlight on the role of AMPK in the larval metamorphosis of M. coruscus, providing a new perspective for the artificial propagation of mollusks.

2. Materials and Methods

2.1. Materials

All mussel larvae were produced in Gouqi Island, Zhoushan, China (122°44′ E; 30°73′ N). The culture method of mussel larvae was modified according to previous studies [26]. After temporary rearing for 7 days under laboratory conditions (18 °C, salinity: 30‰), the larvae were collected for RNA extraction and larval metamorphosis. All extracted RNA was kept at −80 °C after being quickly frozen with liquid nitrogen for the following experiments.

2.2. cDNA Full-Length Cloning and Identification of Three McAMPK Subunits

The total RNA was obtained from the tested larvae utilizing an RNAiso Plus kit (TaKaRa, Shiga, Japan) following the manufacturer’s instructions. The first strand cDNA synthesis for quantitative real-time PCR (qRT-PCR) was operated with a PrimeScriptTM RT reagent kit with gDNA Eraser (Takara, Dalian, China). Both 3′ and 5′ cDNA for rapid amplification of cDNA ends (RACE) PCR were collected with a SMARTer™ RACE 5′/3′ Kit (Clontech, Mountain View, CA, USA). 5′-RACE and 3′-RACE were accomplished using specific primers that were paired with general primers, respectively (Table 1). The touchdown PCR and nested PCR of 5′-RACE and 3′-RACE and the subsequent experimental operation were performed based on the method of Zhu and Li [10,27].

2.3. Sequence Analysis of McAMPK Genes

The amino acid sequences were predicted utilizing ORF finder (http://www.ncbi.nlm.nih.gov/gorf/gorf.html (accessed on 14 May 2021)). Expasy (http://www.expasy.org/tools/protparam.html (accessed on 14 May 2021)) was utilized to predict the molecular weight and theoretical pI of protein. The protein domains were predicted by SMART (http://smart.embl-heidelberg.de/ (accessed on 14 May 2021)). SOPMA (https://npsa-prabi.ibcp.fr (accessed on 14 May 2021)) and SWISS-MODEL (https://www.swissmodel.expasy.org/ (accessed on 7 November 2022)) online programs were used to predict McAMPK secondary and tertiary structures, respectively. The predicted amino acid sequences of McAMPKs aligned with other organisms were performed by DNAman (Version 6). MEGA 7.0 was applied to build phylogenetic trees using the maximum likelihood method [28], and the bootstrap replicates were set to 1000 [29].

2.4. Pharmacologic Experiment of Larval Metamorphosis

In this study, three AMPK activators, Adenosine monophosphate (AMP for short), Metformin (MET for short), and Acadesine (AICAR for short) and one AMPK inhibitor Compound C (CC for short) were used to treat mussel larvae (Table 2). The induction tests of larval metamorphosis by AMP, MET, and AICAR were performed using the strategies outlined in earlier research, and the tested larvae were continuously immersed in a test solution of activators during the entire test [18,30].
Each group of 20 larvae was immersed in a glass Petri dish (Ø 60 mm × 19 mm height) with autoclaved filtered seawater (AFSW), and the chemical solution and the total volume was 20 mL. Each test group—Blank (AFSW), EPI (epinephrine, 10−4 M), AMP, MET, and AICAR—was set up with 6 replicates. Blank was set as a blank control, and EPI was a positive control. EPI-inducted tests on larvae referred to the procedures represented by Yang [38]. After being treated with the working solutions of AMPK antagonist—CC for 2 h—the pediveligers were washed in a glass dish containing AFSW to remove the surface reagent. Then, the larvae were exposed to the AMP solution continuously. The group was named CC+AMP. The preparation of chemical stock solutions and working solutions was based on previous methods [30].

2.5. RNAi

The small interfering RNA (siRNA) sequences designed for McAMPKs were all synthesized by GenePharma Co. Ltd. (Shanghai, China). The confirmation of the negative control sequence (nonsense siRNA) referred to a previous study [27]. The siRNA sequences are shown in Table 3.
The siRNA was transfected into M. coruscus pediveligers by electroporation using the procedures published in a previous study [27]. Briefly, about 200 pediveliger larvae were washed with AFSW before being transferred into a 1.5 mL centrifuge tube (Rnase-free, Dnase-free), which contained 1 mL AFSW (as a blank control) or 1 mL AFSW with 1.2 μg nonsense siRNA (as a negative control group) or 1 mL AFSW with 1.2 μg AMPKα siRNA, 1.2 μg AMPKβ siRNA, or 0.8 μg AMPKγ siRNA (as experimental groups), respectively; then, the samples were incubated for 5 min. Subsequently, the mixtures of siRNA and larvae were sucked from the centrifuge tube into the electroporation reaction cup, and the samples were electroporated with square-wave pulses using Gene Pulser Xcell (Bio-Rad, Hercules, CA, USA). Next, the electroporated mixtures of siRNA and larvae were kept at 25 °C for 10 min; then, they were transferred to the fresh AFSW for 48 h. Finally, the larvae were collected for extracting total RNA and detecting the knock-down effect of the target genes (McAMPKα, McAMPKβ, and McAMPKγ) by qRT-PCR. The experiment was repeated independently 4 times. The experimental groups are shown in Table 4.

2.6. Larval Metamorphosis and RNAi

The RNAi experiments were performed as described above. After being electroporated, the experimental groups were treated with AMP to observe the larval metamorphosis. In this test, the metamorphosis bioassay was divided into nine groups: the experimental groups with electroporation siRNA (siNC+AMP, siAMPKα+AMP, siAMPKβ+AMP, and siAMPKγ+AMP) and the control groups without siRNA (EPI, Blank, Control, AMP, and Control+AMP). Six replicates were set up for each group, and each replicate cultured 20 electroporated larvae. During this period, the metamorphosis and survival rate of larvae at 0, 24, 48, 72, and 96 h were observed and recorded under an Olympus stereoscopic microscope.

2.7. qRT-PCR Analysis of RNAi

For qRT-PCR, the total RNA of tested larvae was isolated and synthesized cDNA. The primer sequences designed for McAMPK genes and reference gene (EF-1α) are presented in Table 1 [6]. In qRT-PCR analysis, 3 biological repeats and 3 technical repeats were set up in each treatment group.

2.8. Statistical Analysis

The experimental data were pre-processed in Excel 2013 and analyzed by one-way analysis of variance (ANOVA) via SPSS 25.0. Significant differences were symbolized by different lowercase letters (p < 0.05).

3. Results

3.1. Characterization of McAMPK Genes Subsection

The full-length cDNA sequences of three McAMPK genes—McAMPKα, McAMPKβ, and McAMPKγ—were cloned using RACE. The McAMPKα gene was 2079 bp long, which consists of a 1335 bp ORF, a 498 bp 5′-UTR, and a 246 bp 3′-UTR (GenBank accession number: ON310500). The ORF encoded 444 amino acids with a predicted molecular weight of 50.81 kDa and a theoretical pI of 5.55 (Figure 1). The SMART program results showed that McAMPKα possessed a S-TKc domain (KD), which is a catalytic domain and an adenylate sensor. The McAMPKβ gene was 1602 bp long, which consists of a 780 bp ORF, a 354 bp 5′-UTR, and a 468 bp 3′-UTR (GenBank accession number: ON310501). The ORF encoded 259 amino acids with a predicted molecular weight of 28.98 kDa and a theoretical pI of 7.84 (Figure 2). The SMART program results showed that McAMPKβ possessed a carbohydrate-binding module (CBM) region and AMPKBI region. The McAMPKγ gene was 2569 bp long, which consists of an 1872 bp ORF, a 306 bp 5′-UTR, and a 391 bp 3′-UTR (GenBank accession number: ON310502). The ORF encoded 623 amino acids with a predicted molecular weight of 70.34 kDa and a theoretical pI of 9.12 (Figure 3). The SMART program results showed that McAMPKγ possessed the conserved domains, four cystathionine β-synthase (CBS) regions. The predicted isoelectric points of the McAMPK genes were 5.55, 7.84, and 9.12, respectively.
The predicted secondary and tertiary structures of the McAMPKα, McAMPKβ, and McAMPKγ proteins are shown in Figure 4. McAMPKα, McAMPKβ, and McAMPKγ proteins consisted of 20, 3, and 22 α helixes; 9, 16, and 16 β strands; and 29, 20, and 35 coils, respectively.

3.2. Multiple Sequence Alignments and Phylogenetic Analysis of McAMPKs

The predicted McAMPKα protein shares a high sequence identity with Mytilus galloprovincialis (76.26% similarity), Pecten maximus (66.10% similarity), Azumapecten farreri (66.10% similarity), and Crassostrea gigas (64.25% similarity) (Figure 5).
The predicted McAMPKβ protein shares a high sequence identity with Pecten maximus (68.08% similarity), Mizuhopecten yessoensis (68.08% similarity), Azumapecten farreri (67.31% similarity), and Crassostrea gigas (67.16% similarity) (Figure 6).
The predicted McAMPKγ protein shares a high sequence identity with Mytilus galloprovincialis (98.24% similarity), Pecten maximus (64.55% similarity), Mizuhopecten yessoensis (63.48% similarity), and Crassostrea gigas (59.24% similarity) (Figure 7).
A phylogenetic tree was formed using AMPK protein sequences found in M. coruscus, other shellfish, insects, batrachian, and mammals (Figure 8). The phylogenetic tree presents three major clusters in accordance with the three various subunits of AMPK, which provides firm phylogenetic evidence for the identities of the McAMPK genes. McAMPKs have a close evolutionary relationship with other bivalves but are far away from insects and vertebrates.

3.3. AMPK Inhibitor and Activators Regulate M. Coruscus Larval Metamorphosis

To investigate the function of AMPK, three AMPK activators, such as AMP, MET, and AICAR, as well as the inhibitor, CC (5 × 10−5 mM), were utilized to test pediveligers, and the metamorphosis rate was compared with the blank control (Blank) and the positive control (EPI) (Figure 9 and Figure 10A). A total of 0.5 mM AMP had good induction for pediveligers to undergo metamorphosis, whose inductivity was significantly higher than the group Blank and the other two activators (p < 0.05) and had a higher survival rate (Supplementary Figure S1). The post-larvae rate of group CC+AMP treated with CC for 2 h and then treated with AMP for 72 h was significantly lower than the group AMP, which was only treated with AMP (p < 0.05, Figure 10A), and the larval survival rates among the groups were unaffected by the different compound treatments (p > 0.05, Figure 10B).

3.4. Validation of RNAi Effect of McAMPK Genes

To detect the effect of RNAi, the gene expression of the McAMPKs of the electroporated larvae and control larvae were analyzed with qRT-PCR technology. The expression of the reference gene EF-1α was not significantly varied among the different groups (Supplementary Figure S2). It was discovered that the expression of the McAMPKα gene was ablated in the electroporated larvae exposed to McAMPKα siRNA in contrast with the blank control without siRNA (Blank) and the negative control treated with NC siRNA (siNC, p < 0.05, Figure 11A), as well as McAMPKβ and McAMPKγ (p < 0.05, Figure 11B,C). Meanwhile, after 96 h, no significant difference was found in the survival rate of larvae among the groups with siRNA transfection and the blank group (p > 0.05, Figure 11D). The above experimental results showed that electroporation can effectively inhibit the higher expression level of McAMPK genes without affecting the survival status, so it can be used for the subsequent function research of McAMPK genes.

3.5. siRNA Transfection of McAMPK Genes Inhibits M. Coruscus Larval Metamorphosis

In the metamorphosis assays, the pediveliger larvae only with electroporation (Control) were unable to complete metamorphosis, as well as Blank (Figure 12). The post-larvae of the pediveliger larvae that were only being treated with EPI or AMP were 42.24% ± 2.36 and 28.92% ± 1.81, respectively. The groups that were electroporated and AMP-induced (Control+AMP) and electroporated with NC siRNA and AMP-induced (siNC+AMP) were 29.99% ± 1.47 and 28.55% ± 2.36, respectively, which were slightly decreased compared with group EPI. The post-larvae of the pediveliger larvae that were treated with AMP coupled with electroporation with McAMPKα siRNA (siAMPKα+AMP), McAMPKβ (siAMPKβ+AMP), and McAMPKγ siRNA (siAMPKγ+AMP) significantly inhibited larval metamorphosis (p < 0.05, Figure 12) compared with group AMP, Control+AMP, and siNC+AMP. The above results showed that the knock-down of McAMPK genes by RNAi can significantly reduce the metamorphosis rate of pediveliger larvae.

4. Discussion

Metamorphosis is an influential process for the growth and development of many shellfish larvae, which is closely related to the survival of larvae and the economic benefits of aquaculture. In the process of metamorphosis, various physiological changes occur in the organism, which requires much energy. As a master cellular energy regulator, AMPK is essential for organism growth, development, autophagy, and metabolism [14,16]. It is well known that AMPK is essential in mammalian development, and some progress has also been made in insect development and metamorphosis [39,40,41]. Nonetheless, it is still uncertain whether AMPK serves a comparable function in marine invertebrates. In this study, the full length and molecular characteristics of the AMPK genes of M. coruscus were cloned and identified. In addition, through pharmacological and RNAi experiments, it was found that the inhibition or activation of the AMPK signal could reduce or promote the metamorphosis of larvae, respectively, which preliminarily proved that AMPK is fundamental in larval metamorphosis.
The unique AMPKα, AMPKβ, and AMPKγ transcript sequences were obtained by comparing the internal M. coruscus transcriptome data with the NCBI database. Our study cloned three AMPK subunits of M. coruscus, namely, McAMPKα, McAMPKβ, and McAMPKγ. The molecular characteristics and the potential function in the metamorphosis of M. coruscus were analyzed and explored. The McAMPKα protein was predicted to comprise two major conserved domains: the catalytic domain S_TKc and adenylate sensor. Furthermore, the conserved sequence serves as the core structure in AMPKα subunits and is phosphorylated by an upstream kinase (LKB1, CaMKIK) [42,43,44]. The predicted McAMPKβ protein contained a CBM domain and an interaction domain AMPKBI. The McAMPKγ protein is comprised of four CBS domains. The results of the predicted domains were consistent with Scophthalmus maximus [45], Ruditapes philippinarum [46], Ctenopharyngodon idellus [47], and so on. The alignments of the predicted amino acid sequences and the tertiary structures of the three subunits of McAMPKs share a high identity with other species, and the phylogenetic tree presented McAMPK as conserved cross from mollusk to vertebrates, which suggests that AMPK proteins were evolutionarily conserved [46,48].
As a molecular hub for cellular metabolic control, AMPK is essential for controlling organism growth and development [22,39,49]. For instance, AMPK could promote protein phosphatase 2A (PP2A) during Bombyx mori and D. melanogaster metamorphosis in response to high amounts of 20-hydroxyecdysone (20E), resulting in limiting growth speed and body weight [39]. When AMPK signaling is triggered, the synthesis of 20E in Hyphantria cunea larvae is blocked, which leads to the delay of larval metamorphosis [50]. Thus, to explore the involvement of AMPK in the metamorphosis of marine invertebrates, the inhibitor or activator of AMPK was used in the metamorphosis experiment of M. coruscus, and the AMPK genes were knocked down with siRNAs. As an upstream signal, AMP is the key signal to activate AMPK [43,51]. As a recognized inhibitor of AMPK, Compound C is widely used in AMPK studies [39,49,52], and it has been demonstrated to be efficient in inhibiting AMPK signaling in oysters [29]. In this study, the results showed that AMP can dramatically promote the metamorphosis of pediveligers, while CC can significantly decrease the induction activity of AMP on larval metamorphosis. According to our RNAi test, the decrease in AMPK gene expression inhibited the pediveligers to metamorphose into plantigrades. These results are consistent with previous studies [23,52,53]. This suggests that AMPK is required for the larval metamorphosis of mussels, but its specific mechanism in larval metamorphosis is still unclear. Furthermore, AMPK has been studied well in insect metamorphosis, and it has been found that AMPK is involved in regulating the energy metabolism during the metamorphosis of D. melanogaster and B. mori [39]. So, during the larval metamorphosis of marine invertebrates, whether there is a molecular mechanism similar to the AMPK signal of insects to regulate dynamic energy metabolism remains to be further studied.

5. Conclusions

In conclusion, three AMPK genes (AMPKα, AMPKβ, and AMPKγ) in M. coruscus were identified. The above findings demonstrated that AMPK plays a prominent role in the larval metamorphosis of M. coruscus. Our study could be an asset to elucidate the molecular mechanisms in the metamorphosis of economical marine shellfish, which can be valuable for mussel aquaculture.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes13122384/s1, Figure S1: The survival rate of M. coruscus larvae at 72 h after being treated with AMPK activators. Figure S2: The expression of EF-1α in different groups.

Author Contributions

J.-L.Y.: supervision, resources, funding acquisition, project administration, and writing—review and editing. X.L.: experimental design and writing—review and editing. W.Z. and Y.W.: investigation, experimental operation, data analysis, and writing—original manuscript. X.H., Z.Z. and Y.Z.: sample collection and processing. All authors contributed to the article and approved the manuscript.

Funding

This study was supported by the National Key Research and Development Program of China (2020YFD0900804), the Program of Shanghai Academic Research Leader (20XD1421800), the Science and technology innovation action plan “The Belt and Road” international joint laboratory construction projects, Science and Technology Commission of Shanghai (19590750500), and the National Natural Science Foundation of China (41876159).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, J.; Feng, D.; Liu, J.; Xu, J.; Chen, K.; Li, Y.; Zhu, Y.; Liang, X.; Lu, Y. Chromosome-level genome assembly of the hard-shelled mussel Mytilus coruscus, a widely distributed species from the temperate areas of East Asia. Gigascience 2021, 10, giab024. [Google Scholar] [CrossRef] [PubMed]
  2. Yang, J.; Li, S.; Li, Y.; Liu, Z.; Liang, X.; Bao, W.; Liu, J. Effects of neuroactive compounds, ions and organic solvents on larval metamorphosis of the mussel Mytilus coruscus. Aquaculture 2013, 396–399, 106–112. [Google Scholar] [CrossRef]
  3. Bishop, C.D.; Erezyilmaz, D.F.; Flatt, T.; Georgiou, C.D.; Hadfield, M.G.; Heyland, A.; Hodin, J.; Jacobs, M.W.; Maslakova, S.A.; Pires, A.; et al. What is metamorphosis? Integr. Comp. Biol. 2006, 46, 655–661. [Google Scholar] [CrossRef] [PubMed]
  4. Ishizuya-Oka, A. Amphibian organ remodeling during metamorphosis: Insight into thyroid hormone-induced apoptosis. Dev. Growth Differ. 2011, 53, 202–212. [Google Scholar] [CrossRef] [PubMed]
  5. McMenamin, S.K.; Parichy, D.M. Metamorphosis in teleosts. Curr. Top. Dev. Biol. 2013, 103, 127–165. [Google Scholar] [PubMed] [Green Version]
  6. Liang, X.; Chen, K.; Li, Y.; Bao, W.; Yoshida, A.; Osatomi, K.; Yang, J. An α2-adrenergic receptor is involved in larval metamorphosis in the mussel, Mytilus coruscus. Biofouling 2019, 35, 986–996. [Google Scholar] [CrossRef] [PubMed]
  7. Espinel-Velasco, N.; Aguera, A.; Lamare, M. Sea urchin larvae show resilience to ocean acidification at the time of settlement and metamorphosis. Mar. Environ. Res. 2020, 159, 104977. [Google Scholar] [CrossRef]
  8. Dubrovsky, E.B. Hormonal cross talk in insect development. Trends Endocrinol. Metab. 2005, 16, 6–11. [Google Scholar] [CrossRef]
  9. Joyce, A.; Vogeler, S. Molluscan bivalve settlement and metamorphosis: Neuroendocrine inducers and morphogenetic responses. Aquaculture 2018, 487, 64–82. [Google Scholar] [CrossRef]
  10. Zhu, Y.; Zhang, Y.; Liu, Y.; Li, Y.; Yoshida, A.; Osatomi, K.; Yang, J.; Liang, X. Nitric oxide negatively regulates larval metamorphosis in hard-shelled mussel (Mytilus coruscus). Front. Mar. Sci. 2020, 7, 356. [Google Scholar] [CrossRef]
  11. He, J.; Wu, Z.; Chen, L.; Dai, Q.; Hao, H.; Su, P.; Ke, C.; Feng, D. Adenosine triggers larval settlement and metamorphosis in the mussel Mytilopsis sallei through the ADK-AMPK-FoxO pathway. ACS Chem. Biol. 2021, 16, 1390–1400. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hoyer-Hansen, M.; Bastholm, L.; Szyniarowski, P.; Campanella, M.; Szabadkai, G.; Farkas, T.; Bianchi, K.; Fehrenbacher, N.; Elling, F.; Rizzuto, R.; et al. Control of macroautophagy by calcium, calmodulin-dependent kinase kinase-β, and Bcl-2. Mol. Cell. 2007, 25, 193–205. [Google Scholar] [CrossRef] [PubMed]
  14. Herzig, S.; Shaw, R.J. AMPK: Guardian of metabolism and mitochondrial homeostasis. Nat. Rev. Mol. Cell Biol. 2018, 19, 121–135. [Google Scholar] [CrossRef] [Green Version]
  15. Hardie, D.G.; Ross, F.A.; Hawley, S.A. AMPK: A nutrient and energy sensor that maintains energy homeostasis. Nat. Rev. Mol. Cell Biol. 2012, 13, 251–262. [Google Scholar] [CrossRef] [Green Version]
  16. Mihaylova, M.M.; Shaw, R.J. The AMPK signalling pathway coordinates cell growth, autophagy and metabolism. Nat. Cell Biol. 2011, 13, 1016–1023. [Google Scholar] [CrossRef]
  17. Hindupur, S.K.; Gonzalez, A.; Hall, M.N. The opposing actions of target of rapamycin and AMP-activated protein kinase in cell growth control. Cold Spring Harb. Perspect. Biol. 2015, 7, a019141. [Google Scholar] [CrossRef] [Green Version]
  18. Bultot, L.; Guigas, B.; Von Wilamowitz-Moellendorff, A.; Maisin, L.; Vertommen, D.; Hussain, N.; Beullens, M.; Guinovart, J.J.; Foretz, M.; Viollet, B.; et al. AMP-activated protein kinase phosphorylates and inactivates liver glycogen synthase. Biochem. J. 2012, 443, 193–203. [Google Scholar] [CrossRef] [Green Version]
  19. Ke, R.; Xu, Q.; Li, C.; Luo, L.; Huang, D. Mechanisms of AMPK in the maintenance of ATP balance during energy metabolism. Cell Biol. Int. 2018, 42, 384–392. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, J.; Xu, Q. Influence of soybean meal on intestinal mucosa metabolome and effects of adenosine monophosphate-activated protein kinase signaling pathway in mirror carp (Cyprinus carpio Songpu). Front. Mar. Sci. 2022, 9, 844716. [Google Scholar] [CrossRef]
  21. Chan, A.Y.; Dyck, J.R. Activation of AMP-activated protein kinase (AMPK) inhibits protein synthesis: A potential strategy to prevent the development of cardiac hypertrophy. Can. J. Physiol. Pharmacol. 2005, 83, 24–28. [Google Scholar] [CrossRef] [PubMed]
  22. Rashid, S.; Wong, C.; Roy, R. Developmental plasticity and the response to nutrient stress in Caenorhabditis elegans. Dev. Biol. 2021, 475, 265–276. [Google Scholar] [CrossRef] [PubMed]
  23. Marzano, M.; Herzmann, S.; Elsbroek, L.; Sanal, N.; Tarbashevich, K.; Raz, E.; Krahn, M.P.; Rumpf, S. AMPK adapts metabolism to developmental energy requirement during dendrite pruning in Drosophila. Cell Rep. 2021, 37, 110024. [Google Scholar] [CrossRef] [PubMed]
  24. Videla, J.A.; Chaparro, O.R.; Thompson, R.J.; Concha, I.I. Role of biochemical energy reserves in the metamorphosis and early juvenile development of the oyster Ostrea chilensis. Mar. Biol. 1998, 132, 635–640. [Google Scholar] [CrossRef]
  25. Chen, Z.; Zhang, H.; Wang, H.; Matsumura, K.; Wong, Y.; Ravasi, T.; Qian, P. Quantitative proteomics study of larval settlement in the Barnacle Balanus amphitrite. PLoS ONE 2014, 9, e88744. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, C.; Bao, W.; Gu, Z.; Li, Y.; Liang, X.; Ling, Y.; Cai, S.; Shen, H.; Yang, J. Larval settlement and metamorphosis of the mussel Mytilus coruscus in response to natural biofilms. Biofouling 2012, 28, 249–256. [Google Scholar] [CrossRef]
  27. Li, Y.; Cheng, Y.; Chen, K.; Cheng, Z.; Zhu, X.; Cardoso, J.C.R.; Liang, X.; Zhu, Y.; Power, D.M.; Yang, J. Thyroid hormone receptor: A new player in epinephrine-induced larval metamorphosis of the hard-shelled mussel. Gen. Comp. Endocrinol. 2020, 287, 113347. [Google Scholar] [CrossRef]
  28. 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] [Green Version]
  29. He, Z.; Mao, F.; Lin, Y.; Li, J.; Zhang, X.; Zhang, Y.; Xiang, Z.; Noor, Z.; Zhang, Y.; Yu, Z. Molecular characteristics of AMPK and its role in regulating the phagocytosis of oyster hemocytes. Fish Shellfish Immunol. 2019, 93, 416–427. [Google Scholar] [CrossRef]
  30. Yang, J.; Satuito, C.G.; Bao, W.; Kitamura, H. Induction of metamorphosis of pediveliger larvae of the mussel Mytilus galloprovincialis Lamarck, 1819 using neuroactive compounds, KCl, NH4Cl and organic solvents. Biofouling 2008, 24, 461–470. [Google Scholar] [CrossRef]
  31. Zhou, G.; Myers, R.; Li, Y.; Chen, Y.; Shen, X.; Fenyk-Melody, J.; Wu, M.; Ventre, J.; Doebber, T.; Fujii, N.; et al. Role of AMP-activated protein kinase in mechanism of metformin action. J. Clin. Investig. 2001, 108, 1167–1174. [Google Scholar] [CrossRef] [PubMed]
  32. Kim, J.; Yang, G.; Kim, Y.; Kim, J.; Ha, J. AMPK activators: Mechanisms of action and physiological activities. Exp. Mol. Med. 2016, 48, e224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chen, L.; Jiao, Z.; Zheng, L.; Zhang, Y.; Xie, S.; Wang, Z.; Wu, J. Structural insight into the autoinhibition mechanism of AMP-activated protein kinase. Nature 2009, 459, 1146–1149. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, Y.; Wang, Y.; Bao, C.; Xu, Y.; Shen, H.; Chen, J.; Yan, J.; Chen, Y. Metformin interacts with AMPK through binding to γ subunit. Mol. Cell Biochem. 2012, 368, 69–76. [Google Scholar] [CrossRef] [PubMed]
  35. Lee, H.; Zandkarimi, F.; Zhang, Y.; Meena, J.K.; Kim, J.; Zhuang, L.; Tyagi, S.; Ma, L.; Westbrook, T.F.; Steinberg, G.R.; et al. Energy-stress-mediated AMPK activation inhibits ferroptosis. Nat. Cell Biol. 2020, 22, 225–234. [Google Scholar] [CrossRef] [PubMed]
  36. Rao, E.; Zhang, Y.; Li, Q.; Hao, J.; Egilmez, N.K.; Suttles, J.; Li, B. AMPK-dependent and independent effects of AICAR and compound C on T-cell responses. Oncotarget 2016, 7, 33783–33795. [Google Scholar] [CrossRef] [Green Version]
  37. Saito, S.; Furuno, A.; Sakurai, J.; Park, H.R.; Shin-Ya, K.; Tomida, A. Compound C prevents the unfolded protein response during glucose deprivation through a mechanism independent of AMPK and BMP signaling. PLoS ONE 2012, 7, e45845. [Google Scholar] [CrossRef]
  38. Yang, J.; Li, Y.; Bao, W.; Satuito, C.G.; Kitamura, H. Larval metamorphosis of the mussel Mytilus galloprovincialis Lamarck, 1819 in response to neurotransmitter blockers and tetraethylammonium. Biofouling 2011, 27, 193–199. [Google Scholar] [CrossRef] [Green Version]
  39. Yuan, D.; Zhou, S.; Liu, S.; Li, K.; Zhao, H.; Long, S.; Liu, H.; Xie, Y.; Su, Y.; Yu, F.; et al. The AMPK-PP2A axis in insect fat body is activated by 20-hydroxyecdysone to antagonize insulin/IGF signaling and restrict growth rate. Proc. Natl. Acad. Sci. USA 2020, 117, 9292–9301. [Google Scholar] [CrossRef]
  40. Evans, J.J.; Xiao, C.; Robertson, R.M. AMP-activated protein kinase protects against anoxia 2 in Drosophila melanogaster. PLoS ONE 2017, 8, 30–39. [Google Scholar] [CrossRef]
  41. O’Neill, H.M.; Maarbjerg, S.J.; Crane, J.D.; Jeppesen, J.; Jorgensen, S.B.; Schertzer, J.D.; Shyroka, O.; Kiens, B.; van Denderen, B.J.; Tarnopolsky, M.A.; et al. AMP-activated protein kinase (AMPK) β1β2 muscle null mice reveal an essential role for AMPK in maintaining mitochondrial content and glucose uptake during exercise. Proc. Natl. Acad. Sci. USA 2011, 108, 16092–16097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Woods, A.; Johnstone, S.R.; Dickerson, K.; Leiper, F.C.; Fryer, L.G.; Neumann, D.; Schlattner, U.; Wallimann, T.; Carlson, M.; Carling, D. LKB1 is the upstream kinase in the AMP-activated protein kinase cascade. Curr. Biol. 2003, 13, 2004–2008. [Google Scholar] [CrossRef] [Green Version]
  43. Hawley, S.A.; Selbert, M.A.; Goldstein, E.G.; Edelman, A.M.; Carling, D.; Hardie, D.G. 5′-AMP activates the AMP-activated protein kinase cascade, and Ca2+/calmodulin activates the calmodulin-dependent protein kinase I cascade, via three independent mechanisms. J. Biol. Chem. 1995, 270, 27186–27191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Hawley, S.A.; Boudeau, J.; Reid, J.L.; Mustard, K.J.; Udd, L.; Mäkelä, T.P.; Alessi, D.R.; Hardie, D.G. Complexes between the LKB1 tumor suppressor, STRAD α/β and MO25 α/β are upstream kinases in the AMP-activated protein kinase cascade. J. Biol. 2003, 2, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zeng, L.; Liu, B.; Wu, C.; Lei, J.; Xu, M.; Zhu, A.; Zhang, J.; Hong, W. Molecular characterization and expression analysis of AMPK α subunit isoform genes from Scophthalmus maximus responding to salinity stress. Fish Physiol. Biochem. 2016, 42, 1595–1607. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, J.; Fang, L.; Wu, Q.; Li, D.; Huo, Z.; Yan, X. Genome-wide identification and characterization of the AMPK genes and their distinct expression patterns in response to air exposure in the Manila clam (Ruditapes philippinarum). Genes Genom. 2020, 42, 1–12. [Google Scholar] [CrossRef]
  47. Wu, W.; Sun, J.; Ji, H.; Yu, H.; Zhou, J. AMP-activated protein kinase in the grass carp Ctenopharyngodon idellus: Molecular characterization, tissue distribution and mRNA expression in response to overwinter starvation stress. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2020, 246–247, 110457. [Google Scholar] [CrossRef]
  48. Sanz, P. AMP-activated protein kinase: Structure and regulation. Curr. Protein Pept. Sci. 2008, 9, 478–492. [Google Scholar] [CrossRef]
  49. Liu, X.; Chhipa, R.R.; Nakano, I.; Dasgupta, B. The AMPK inhibitor compound C is a potent AMPK-independent antiglioma agent. Mol. Cancer Ther. 2014, 13, 596–605. [Google Scholar] [CrossRef] [Green Version]
  50. Zou, H.; Zou, H.; Li, X.; Qiu, Q.; Geng, N.; Zhang, B.; Yan, G.; Zhang, Z.; Zhang, S.; Yao, B.; et al. Metformin-induced AMPK activation suppresses larval growth and molting probably by disrupting 20E synthesis and glycometabolism in fall webworm, Hyphantria cunea Drury. Pestic. Biochem. Physiol. 2022, 183, 105083. [Google Scholar] [CrossRef]
  51. Lin, S.; Hardie, D.G. AMPK: Sensing glucose as well as cellular energy status. Cell Metab. 2018, 27, 299–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lee, J.H.; Koh, H.; Kim, M.; Kim, Y.; Lee, S.Y.; Karess, R.E.; Lee, S.H.; Shong, M.; Kim, J.M.; Kim, J.; et al. Energy-dependent regulation of cell structure by AMP-activated protein kinase. Nature 2007, 447, 1017–1020. [Google Scholar] [CrossRef] [PubMed]
  53. Lippai, M.; Csikos, G.; Maroy, P.; Lukacsovich, T.; Juhasz, G.; Sass, M. SNF4Aγ, the Drosophila AMPK γ subunit is required for regulation of developmental and stress-induced autophagy. Autophagy 2008, 4, 476–486. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The nucleotide and predicted protein sequences of McAMPKα. The uppercase letters in cDNA represent the open reading frame (ORF), and the lowercase letters represent the 5′ and 3′ noncoding regions; * indicates stop codon. The start (ATG) and stop (TAA) codons are presented in bold and underlined. The same as below. Only underlined areas represent the S_TKc domain, and the black-framed regions represent the adenylate sensor domain.
Figure 1. The nucleotide and predicted protein sequences of McAMPKα. The uppercase letters in cDNA represent the open reading frame (ORF), and the lowercase letters represent the 5′ and 3′ noncoding regions; * indicates stop codon. The start (ATG) and stop (TAA) codons are presented in bold and underlined. The same as below. Only underlined areas represent the S_TKc domain, and the black-framed regions represent the adenylate sensor domain.
Genes 13 02384 g001
Figure 2. The nucleotide and predicted protein sequences of McAMPKβ. The underlined areas represent the CBM domain, and the black-framed regions represent the interaction domain, AMPKBI.
Figure 2. The nucleotide and predicted protein sequences of McAMPKβ. The underlined areas represent the CBM domain, and the black-framed regions represent the interaction domain, AMPKBI.
Genes 13 02384 g002
Figure 3. The nucleotide and predicted protein sequences of McAMPKγ. The black-framed region represents the low complexity domain, and the underlined areas represent the four CBS domain repeats.
Figure 3. The nucleotide and predicted protein sequences of McAMPKγ. The black-framed region represents the low complexity domain, and the underlined areas represent the four CBS domain repeats.
Genes 13 02384 g003
Figure 4. The predicted secondary and tertiary structure of McAMPK.
Figure 4. The predicted secondary and tertiary structure of McAMPK.
Genes 13 02384 g004
Figure 5. The multiple sequence alignment of McAMPKα with other mollusks. The dark blue area: complete similarity; the pink area: 75% similarity; the blue area: 50% similarity; and the yellow area: more than 33% similarity. The same as below. The underlined area represents the S_TKc domain. The square-dotted line represents the adenylate sensor domain. The accession number of the AMPKα sequence used: Pencten maximus (XP_033736754.1), Crassostrea gigas (XP_034311308.1), Azumapecten farreri (QFR39800.1), and Mytilus galloprovincialis (VDI56011.1).
Figure 5. The multiple sequence alignment of McAMPKα with other mollusks. The dark blue area: complete similarity; the pink area: 75% similarity; the blue area: 50% similarity; and the yellow area: more than 33% similarity. The same as below. The underlined area represents the S_TKc domain. The square-dotted line represents the adenylate sensor domain. The accession number of the AMPKα sequence used: Pencten maximus (XP_033736754.1), Crassostrea gigas (XP_034311308.1), Azumapecten farreri (QFR39800.1), and Mytilus galloprovincialis (VDI56011.1).
Genes 13 02384 g005
Figure 6. The multiple sequence alignment of McAMPKβ with other mollusks. The double underline area represents the CBM domain. The dashed line area represents the AMPKBI domain. The accession number of the AMPKβ sequence used: Pecten maximus (XP_033735434.1), Crassostrea gigas (XP_011450446.2), Azumapecten farreri (QFR39801.1), and Mizuhopecten yessoensis (XP_021357013.1).
Figure 6. The multiple sequence alignment of McAMPKβ with other mollusks. The double underline area represents the CBM domain. The dashed line area represents the AMPKBI domain. The accession number of the AMPKβ sequence used: Pecten maximus (XP_033735434.1), Crassostrea gigas (XP_011450446.2), Azumapecten farreri (QFR39801.1), and Mizuhopecten yessoensis (XP_021357013.1).
Genes 13 02384 g006
Figure 7. Multiple sequence alignment of McAMPKγ with other mollusks. The double-dashed line area represents the CBS domain; the long-dashed line area represents the low complexity. The accession number of the AMPKγ sequence used: Pecten maximus (XP_033762248.1), Crassostrea gigas (XP_034310765.1), Mizuhopecten yessoensis (XP_021369706.1), and Mytilus galloprovincialis (VDI78590.1).
Figure 7. Multiple sequence alignment of McAMPKγ with other mollusks. The double-dashed line area represents the CBS domain; the long-dashed line area represents the low complexity. The accession number of the AMPKγ sequence used: Pecten maximus (XP_033762248.1), Crassostrea gigas (XP_034310765.1), Mizuhopecten yessoensis (XP_021369706.1), and Mytilus galloprovincialis (VDI78590.1).
Genes 13 02384 g007
Figure 8. A phylogenetic analysis of McAMPKs. Numbers at the tree present the bootstrap values from 1000 replicates.
Figure 8. A phylogenetic analysis of McAMPKs. Numbers at the tree present the bootstrap values from 1000 replicates.
Genes 13 02384 g008
Figure 9. Larval metamorphosis of M. coruscus in response to AMPK activators at 72 h. EPI: epinephrine (10−4 M); Blank: autoclaved filtered seawater (AFSW); AMP: Adenosine monophosphate; MET: Metformin; and AICAR: Acadesine. The abscissa values represent the different test concentrations of the medicines.
Figure 9. Larval metamorphosis of M. coruscus in response to AMPK activators at 72 h. EPI: epinephrine (10−4 M); Blank: autoclaved filtered seawater (AFSW); AMP: Adenosine monophosphate; MET: Metformin; and AICAR: Acadesine. The abscissa values represent the different test concentrations of the medicines.
Genes 13 02384 g009
Figure 10. The larval metamorphosis and survival rate of M. coruscus at 72 h after being treated with AMPK inhibitors. (A) The rate of post-larvae. (B) The survival rate at different times.
Figure 10. The larval metamorphosis and survival rate of M. coruscus at 72 h after being treated with AMPK inhibitors. (A) The rate of post-larvae. (B) The survival rate at different times.
Genes 13 02384 g010
Figure 11. The effects of RNAi on the relative expression of McAMPK genes and the survival rate of M. coruscus. (A) McAMPKα expression in electroporated with siAMPKα and control larvae, (B) McAMPKβ expression in electroporated with siAMPKβ and control larvae, (C) McAMPKγ expression in electroporated with siAMPKγ and control larvae and (D) The survival rate of different groups electroporated with siAMPKα, siAMPKβ and siAMPKγ, respectively.
Figure 11. The effects of RNAi on the relative expression of McAMPK genes and the survival rate of M. coruscus. (A) McAMPKα expression in electroporated with siAMPKα and control larvae, (B) McAMPKβ expression in electroporated with siAMPKβ and control larvae, (C) McAMPKγ expression in electroporated with siAMPKγ and control larvae and (D) The survival rate of different groups electroporated with siAMPKα, siAMPKβ and siAMPKγ, respectively.
Genes 13 02384 g011
Figure 12. The larval metamorphosis of M. coruscus at 72 h after McAMPK RNAi. Control: electroporation; Control + AMP: electroporation + AMP; siRNA + AMP: negative control siRNA + electroporation + AMP; siAMPKα + AMP: McAMPKα siRNA + electroporation + AMP; siAMPKβ + AMP: McAMPKβ siRNA + electroporation + AMP; and siAMPKγ: McAMPKγ siRNA + electroporation + AMP.
Figure 12. The larval metamorphosis of M. coruscus at 72 h after McAMPK RNAi. Control: electroporation; Control + AMP: electroporation + AMP; siRNA + AMP: negative control siRNA + electroporation + AMP; siAMPKα + AMP: McAMPKα siRNA + electroporation + AMP; siAMPKβ + AMP: McAMPKβ siRNA + electroporation + AMP; and siAMPKγ: McAMPKγ siRNA + electroporation + AMP.
Genes 13 02384 g012
Table 1. The sequences of primers used in this study.
Table 1. The sequences of primers used in this study.
Prime Name(5′-3′ Sequence)Comment
ORF-McAMPKα-FTCTTGGTGAAACACTAGGAACTGGORF
ORF-McAMPKα-RTAGGTGGAGTCTGAGATTTGTTGGCORF
ORF-McAMPKβ-FCAGGGAGCTGATATGGTTTAORF
ORF-McAMPKβ-RGAGGAGGAAGAATAGGTGGTORF
ORF-McAMPKγ-FCAGGGTCAACAGGATTACGAGORF
ORF-McAMPKγ-RTTGAACTATCCCAGAGAGGTGORF
RACE-McAMPKα-R1GGAGCAGCATAATTAGGTGAACC5′ RACE
RACE-McAMPKα-R2TCCGTAGGAGTACTTATCACCTGG5′ RACE
RACE-McAMPKα-F1ACTGCCAACAAATCTCAGACTCC3′ RACE
RACE-McAMPKα-F2TGAGGGAATCGTCAGTAGCATC3′ RACE
RACE-McAMPKβ-R1GAGGAAGAATAGGTGGTCCAGTG5′ RACE
RACE-McAMPKβ-R2TATCCTTTCCGCCACCTTCC5′ RACE
RACE-McAMPKβ-F1GCCCACTGTGAACCAACCTTATTAC3′ RACE
RACE-McAMPKγ-R1CGTAATCCTGTTGACCCTGACC5′ RACE
RACE-McAMPKγ-R2AGCAAACGCACTGAGGTCCT5′ RACE
RACE-McAMPKγ-F1ATATAATGGTGTGAGGGCAGC3′ RACE
RACE-McAMPKγ-F2ATCGGCTGTCCCTGTAGTAGA3′ RACE
q-McAMPKα-FCTACAACAGCACAAATCAGAGqRT-PCR
q-McAMPKα-RCATATGGATCACCATTCAACqRT-PCR
q-McAMPKβ-FAGTATTCAATGGGAGGTGGCGqRT-PCR
q-McAMPKβ-RCAGGCTCTGAAGGGTTGTGCTqRT-PCR
q-McAMPKγ-FATAATGGTGTGAGGGCAGCqRT-PCR
q-McAMPKγ-RCCAGGTAGCAATTTTGTGTqRT-PCR
EF--RT-FCACCACGAGTCTCTCCCTGAqRT-PCR
EF--RT-RGCTGTCACCACAGACCATTCCqRT-PCR
Note: “F” indicates forward primer, and “R” indicated reverse primer.
Table 2. Inducers used in the assay, and manufacturers and concentrations of stock and tested solutions.
Table 2. Inducers used in the assay, and manufacturers and concentrations of stock and tested solutions.
InducerFunction on AMPKManufacture and Cat. No.Concentration (mM)
Stock
Solution
Test
Solution
AMPActivator. Binding to AMPK γ to activate AMPK holoenzyme [31,32,33].MCE, HY-A018150.1, 1, 2.5
MetforminActivator. It can directly interact with AMPK by
binding to AMPK γ [34].
MCE, HY-B062750.1, 1, 4
AICARActivator. An adenosine analogue, its function is similar to AMP [35,36].MCE, HY-1341750.1, 0.5, 2
Compound CInhibitor. Suppressed the phosphorylation of AMPK [36,37].MCE, HY-13418A15 × 10−5
Table 3. The sequence of siRNAs.
Table 3. The sequence of siRNAs.
siRNAsiRNA Sequence (5′-3′)Comment
siRNA-AMPKαGCGGUGAACUCUUUGAUUATTRNAi
siRNA-AMPKβGGUGGCGGAAGGAUAUUUATTRNAi
siRNA-AMPKγGCAGAUCAUUCCUUCAAUATTRNAi
siRNA-NCUUCUCCGAACGUGUCACGUTTRNAi
Table 4. The experimental groups list.
Table 4. The experimental groups list.
GroupsiRNA (μg/mL)ElectroporationEPI (10−4 M)AMP (0.5 mM)
Blank0---
EPI0-+-
Control0+--
AMP0--+
Control+AMP0+-+
siNC+AMP1.2+-+
siAMPKα+AMP0.8+-+
siAMPKβ+AMP0.8+-+
siAMPKγ+AMP1.2+-+
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, W.; Wang, Y.; Hu, X.; Zhou, Z.; Zhu, Y.; Liang, X.; Yang, J.-L. AMPK Promotes Larval Metamorphosis of Mytilus coruscus. Genes 2022, 13, 2384. https://doi.org/10.3390/genes13122384

AMA Style

Zhang W, Wang Y, Hu X, Zhou Z, Zhu Y, Liang X, Yang J-L. AMPK Promotes Larval Metamorphosis of Mytilus coruscus. Genes. 2022; 13(12):2384. https://doi.org/10.3390/genes13122384

Chicago/Turabian Style

Zhang, Wen, Yuyi Wang, Xiaomeng Hu, Zhongsheng Zhou, Youting Zhu, Xiao Liang, and Jin-Long Yang. 2022. "AMPK Promotes Larval Metamorphosis of Mytilus coruscus" Genes 13, no. 12: 2384. https://doi.org/10.3390/genes13122384

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