Skip to main content
Acta Pharmaceutica Sinica. B logoLink to Acta Pharmaceutica Sinica. B
. 2022 Jun 11;12(10):3743–3782. doi: 10.1016/j.apsb.2022.06.004

Autophagy and beyond: Unraveling the complexity of UNC-51-like kinase 1 (ULK1) from biological functions to therapeutic implications

Ling Zou a,b,, Minru Liao b,, Yongqi Zhen b,, Shiou Zhu b, Xiya Chen a, Jin Zhang a,b,, Yue Hao a,, Bo Liu b,
PMCID: PMC9532564  PMID: 36213540

Abstract

UNC-51-like kinase 1 (ULK1), as a serine/threonine kinase, is an autophagic initiator in mammals and a homologous protein of autophagy related protein (Atg) 1 in yeast and of UNC-51 in Caenorhabditis elegans. ULK1 is well-known for autophagy activation, which is evolutionarily conserved in protein transport and indispensable to maintain cell homeostasis. As the direct target of energy and nutrition-sensing kinase, ULK1 may contribute to the distribution and utilization of cellular resources in response to metabolism and is closely associated with multiple pathophysiological processes. Moreover, ULK1 has been widely reported to play a crucial role in human diseases, including cancer, neurodegenerative diseases, cardiovascular disease, and infections, and subsequently targeted small-molecule inhibitors or activators are also demonstrated. Interestingly, the non-autophagy function of ULK1 has been emerging, indicating that non-autophagy-relevant ULK1 signaling network is also linked with diseases under some specific contexts. Therefore, in this review, we summarized the structure and functions of ULK1 as an autophagic initiator, with a focus on some new approaches, and further elucidated the key roles of ULK1 in autophagy and non-autophagy. Additionally, we also discussed the relationships between ULK1 and human diseases, as well as illustrated a rapid progress for better understanding of the discovery of more candidate small-molecule drugs targeting ULK1, which will provide a clue on novel ULK1-targeted therapeutics in the future.

Key words: UNC-51-like kinase 1 (ULK1), Autophagy, Non-autophagy, Biological function, Small-molecule drug, ULK1-targeted therapy, Human diseases

Graphical abstract

This review elucidates the serine/threonine kinase ULK1 regulating human diseases through autophagic and non-autophagic pathways, and further summarizes the therapeutic potential of ULK1-targeted compounds (inhibitors/activators/regulators) in different diseases.

Image 1

1. Introduction: A brief overview of UNC-51-like kinase 1 (ULK1)

Autophagy was proposed by Ashford and Porter after discovering the phenomenon of “eating yourself” in cells in 1962. It is an intracellular degradation system through which cytoplasmic material is transported to the lysosome and degraded, and converted into nutrient recycling for the cell1,2. During phases of insufficient nutrient availability, autophagy will be quickly upregulated to provide energy and maintain basic cell functions. Therefore, cells starved of glucose or amino acid supply or under hypoxia rely on autophagy to preserve their survival3,4. In general, autophagy is a kind of mechanism that promotes cellular survival by keeping energy homeostasis and removing toxic proteins, pathogens, and aged or harmful organelles2. This mechanism was discovered by Yoshinori Ohsumi in a mutant Apg1 yeast in 1993, and for this, he won the 2016 Nobel Prize in Physiology or Medicine5. Intriguingly, Apg1, also known as Atg1, was the first protein found to be involved in the initiation of autophagy in yeast5,6. The following year, human homologous protein ULK1(UNC-51 [Caenorhabditis elegans]-like kinase 1) was isolated by Ogura and his colleagues from mutants of C. elegans displaying aberrant axon extension and development, which is encoding a new Ser/Thr kinase7. Since then, pioneering creations about autophagy initiator Atg1/ULK1 were continuously made by numerous labs around the world. In especial, the molecular and biochemical approaches revealed miscellaneous binding partners, upstream regulators/downstream effectors, signaling mediators of Atg1/ULK1, which bring holds promise for the treatment of autophagy-related diseases. For example, in yeast, Atg1 is complexed with Atg13 and Atg17, which are necessary for autophagosome formation. Similarly, in mammals, ULK1 also interacts with mATG13, ATG101, and FIP200 binding proteins to activate autophagy in the form of complexes8, 9, 10. In addition, it was found that ULK1 is not only a downstream effector of mammalian target of rapamycin (mTOR) but also phosphorylated and activated by AMP-activated protein kinase (AMPK)9,11,12. Due to the evolutionary conservation of Atg1/ULK1 in protein trafficking, energy- and nutrient-sensing kinases can directly regulate Atg1/ULK1 to regulate the distribution and utilization of cellular resources to maintain cellular homeostasis in changing complex physiological environments in vivo13. Therefore, ULK1 is necessary for the occurrence and development of diseases, associated with autophagy or non-autophagy pathways. Certainly, ULK1 has been proved to regulate cancer, neurodegenerative disorders, infection, cardiopathy, and other diseases. Further, a sequence of small molecule compounds targeting ULK1 has been revealed good therapeutic efficacy for multiple human diseases. Shaw's lab discovered and synthesized ULK1 kinase inhibitor SBI-0206965 with high selectivity, which inhibited ULK1-mediated phosphorylation to regulate autophagy and cell survival14. In 2017, first ULK1 targeted activator LYN-1604 was designed and discovered by Liu's lab based on computer-aided drug design methods and screened with kinase and antiproliferative activity. Importantly, LYN-1604 exhibited good therapeutic potential for breast cancer by activation of ULK1 and ULK1 complex (Fig. 1)15.

Figure 1.

Figure 1

The timeline of the autophagic initiator ULK1 from biological functions to therapeutic implications.

Thus, in this review, we focused on summarizing the molecular structures and biological functions of ULK1 in autophagy and beyond autophagy. Besides, we clarified the complex and elaborate protein–protein interaction (PPI) network in which ULK1 resides, its direct regulators, indirect regulators, and substrates. In addition, we introduced the key signaling pathways and molecular mechanism which contributed to complete the role of ULK1 in different human diseases, associated with its autophagy and non-autophagy functions. Moreover, we further discussed the prospects of drug development targeting ULK1 and novel ULK1- for the future therapeutic applications.

2. Molecular structures of ULK1 and its complex

2.1. The structure of ULK1

ULK1 protein composed of 1051 amino acids has a molecular weight of 112 kDa and consists of an N-terminal kinase domain (KD); a C-terminal domain (CTD) for ULK1 and its complex to interact with; as well as a proline/serine (PS) for the most post-translational modification16,17. For ULK1, KD promotes kinase activity, while CTD contains two microtubule-interacting and transport (MIT) domains that encourage the interaction of ULK1 with mATG13 and FIP200. The PS region connecting KD and MIT are not well conserved, formed from about 500 amino acids, which is mainly responsible for the interaction of ULK1 with autophagic microtubule-associated protein light chain 3 (LC3) and the phosphorylation modification of ULK1 by upstream kinases (AMPK, etc.). Notably, KD and MIT exhibit high sequence conservation in the Atg1/ULK1 family of kinases, both in yeast and mammals, in comparison to the previously mentioned regions (Fig. 2A)18,19.

Figure 2.

Figure 2

The primary, secondary and tertiary structures of ULK1. (A) Structure and activation of ULK1. (B) Amino acid sequence comparison of Atg1, UNC-51, ULK1, and ULK2. (C) Compositional comparison of the nematode UNC-51 complex, the yeast Atg1 complex, and the mammalian ULK1 complex.

ULK1 contains three homologous proteins: yeast Atg1, mammalian ULK2, and C. elegans UNC-51, which show similarities in structure and function20,21. Atg1 was the first Ser/Thr kinase in the Atg protein and is essential for autophagy in yeast22. Atg1 is composed of 893 amino acids and has three regions: an N-terminal Ser/Thr KD; a C-terminal globular domain containing two MIT tandem microtubules; as well as an intrinsically disordered region that function as junctions. In addition, as a pivotal regulatory gene controlling axonal elongation in C. elegans, UNC-51 shares strong homology with Atg117. UNC-51 contains 855 amino acids and is also split into three parts: N-terminal KD, CTD, and spacer domain23. It is also confirmed that UNC-51 is indispensable for neuron development through mediation of vesicular transport in axons. For C. elegans, deficiency of the Unc-51 gene leads to various abnormal accumulation of vesicles and other membranous structures in neuronal cells7,17. Furthermore, ULK2 is also a functional homologue of yeast Atg1. Except for the N-terminal kinase domain, ULK1 and ULK2 have significant homology in their C-terminal region17. However, in mammals, ULK2 is redundant in its autophagic function compared to its paralog ULK124. Fig. 2B illustrates a sequence comparison of ULK1/2, Atg1, and UNC-51. The overall similarity of ULK1 to ULK2, UNC-51, and Atg1 was 55%, 41%, and 29%, respectively7.

2.2. The structure of the ULK1 complex

ULK1 typically functions by forming a tetrameric complex with ATG13, ATG101, and FIP200 to initiate starvation-induced autophagy19. A corresponding analogy is also in yeast, where Atg1 becomes a pentameric complex with Atg13 (the ULK1 and ATG13 homolog), Atg17 (the mammalian FIP200 homolog), Atg29, and Atg31, which are the most upstream autophagy initiation factor19,25. After forming a complex with Atg29 and Atg31, Atg17 further interacts with Atg1 and Atg13 to promote the organization of pre-autophagosome structures upon starvation26. In addition, in C. elegans, EPG-9 (homologous gene of ATG101) directly interacts with EPG-1(homologous gene of mAtg13/Atg13) and EPG-7 and forms a complex in the aggrephagy pathway to initiate autophagy27,28.

ULK1 regulates the initiation of autophagy by forming the ULK1–ATG13–FIP200–ATG101 stable complex. Therein, ATG13 and FIP200 cooperate in the autophagy process29. ATG13 has a very short peptide motif at the C-terminus that connects to ULK1, and it also acts as an adaptor protein to recruit components of the ULK1 complex, including ULK1, FIP200, and ATG101, to the core to initiate autophagy30,31. Meanwhile, residues 582–585 in FIP200 are required to interact with ATG13 to function, and mutation of these residues affects normal autophagy10. As another component of the ULK1 complex, ATG101 is a stabilizer of ATG13 and an auxiliary protein of the complex, which is not regulated by nutritional conditions. It also functions as a bridge to link ULK1 and the downstream PtdIns3K complex during mammalian autophagy32,33. In addition, the ATG101–ATG13 complex was found to be a rigid structure by measuring the crystal structure of yeast ATG101, and ATG101 could stabilize the Atg13 HORMA fold in higher eukaryotes34, 35, 36.

Autophagy initiation complexes are loosely bound in yeast. Because Atg1 first forms a relatively stable sub-complex with Atg13, meanwhile, Atg17, Atg31, and Atg29 also form a stable sub-complex, and finally, these two parts combine with moderate binding force37. Atg17–Atg31–Atg29 complex is a double crescent dimer and migrates to pre-autophagosome, with Atg31 acting as the bridge between the other two proteins. This curved structure helps the complex bind Atg1–Atg13 to ATG-containing vesicles, which may be the source of the membrane formed by the autophagosome38.

The situation in C. elegans is relatively simple. The EPG-9–EPG-1 complex accumulates in C. elegans and then interacts with UNC-51. Among them, EPG-1 is similar to Atg13 in yeast39. EPG-9 could form a complex with EPG-1 by direct interaction in autophagy process in C. elegans27. EPG-7, a protein similar to human FIP200 and yeast Atg11, acts as a scaffold protein but is unnecessary for starvation-induced autophagic degradation28. In addition, in C. elegans, UNC-51 directly binds and affects the activities of VAB-8 and UNC-14, where VAB-8 is involved in axonal backward migration and growth, and UNC-14 regulates UNC-5-related sub-cellular localization and actin 1-dependent synaptic vesicle transport. These two proteins are involved in axon conduction and growth and are independent compared to autophagic proteins EPG-1 and EPG-9 (Fig. 2C)40.

3. The molecular basis underlying the regulation mechanism of ULK1

3.1. Transcriptional and post-transcriptional regulation of ULK1

3.1.1. Transcriptional regulation

Transcriptional regulation of ULK1 is the initial and most critical influencer controlling the gene expression and active protein content of ULK1. In this process, several genes, DNA-binding proteins, and transcription factors play crucial roles (Table 1). Recently, ZFP36L2 and RAB13 have been found to positively regulate the transcription level and protein expression level of ULK1 based on multi-omics approaches41. In addition, transcription factors, such as NFE2L2/NRF242, NF-E243, forkhead box O (FOXO) 144, FOXO345, hepatocyte nuclear factor 4 alpha (Hnf4α)46, activating transcription factor 4 (ATF4)47, signal transducer and activator of transcription-1 (STAT1)48, and DAXX49 have been proved to associated with the transcriptional regulation of ULK1. For instance, FOXO1 and FOXO3 are stress-responsive transcription factors, they could directly enhance the transcription of ULK1 and other Atg genes to initiate autophagy44,45. In nonalcoholic hepatic steatosis, Hnf4α could directly activate the transcription of ULK1 to stimulate autophagy46. ATF4, a known endoplasmic reticulum (ER) stress regulator, could transcriptional up-regulation of ULK1 in response to severe cellular stress, including hypoxia, ER stress and anti-angiogenic therapy47. STAT1 is a transcriptional suppressor of ULK1, which negatively regulates the mRNA and protein levels of ULK148. Likewise, the transcriptional repressor DAXX inhibits the expression of ULK1 in vivo to suppress autophagy, promoting PCa tumorigenicity49. Additionally, cAMP response element-binding protein (CREB) is a new transcriptional activator of autophagy, which upregulates the autophagy gene ULK150. Further, nuclear receptor subfamily 1 group D member 1 (NR1D1) can directly enhance the transcription of Ulk1 during spermatogenesis in mice. Interestingly, STRA8 can inhibit the expression of NR1D1 via binding to the Nr1d1 promoter thus to impede NR1D1-regulated ULK1 up-regulation51. Notably, phospho-p53 (Ser392) could increase ULK1 expression by binding to ULK1 promoter to stimulate autophagy52. High Mobility Group A1 (HMGA1) is an architectural chromatin protein, recently was found to transcriptionally up-regulate ULK1 to facilitate autophagy in cancer cells53. Intriguingly, high mobility group nucleosomal binding domain 2 (HMGN2) and H3K27ac co-regulates the transcription and expression of ULK1. Of them, HMGN2 recruitment at the ULK1 promoter to accelerate ULK1 expression while H3K27ac negatively affect the transcription of ULK154. Moreover, the spliceosome PRPF8 was found to facilitate mitophagy by affecting ULK1 mRNA splicing and might result in retinitis pigmentosa55. PGC-1α, a co-regulator of gene expression in mitochondrial biogenesis, promotes autophagy and mitophagy through ULK1 to reduce NLRP3 inflammasome, suppressing neuroinflammation in acute ischemic stroke56 (Fig. 3A).

Table 1.

Transcriptional and post-transcriptional regulation of ULK1.

No. Name Classification Positive regulation/
negative regulation
Mechanism Ref.
1 ZFP36L2 Gene Positive regulation ZFP36L2 positively regulates the transcription level and protein expression level of ULK1 41
2 RAB13 Gene Positive regulation RAB13 positively regulates the transcription level and protein expression level of ULK1 41
3 NFE2L2/NRF2 Transcription factors Positive regulation NFE2L2/NRF2 activates the expression of ULK1 42
4 NF-E2 Transcription factors Positive regulation NF-E2 increase the expression of ULK1 43
5 FOXO1 Transcription factors Positive regulation FoxO1 promotes the expression of ULK1 44
6 FOXO3 Transcription factors Positive regulation FoxO3 activation increases the abundance of mRNA and protein levels of ULK1 45
7 Hnf4α Transcription factors Positive regulation Hnf4α directly activating transcription of ULK1 46
8 miR-214-3p McroRNA Negative regulation Suppressing ULK1 expression through direct binding at a 3′ untranslated region sequence 46
9 ATF4 Transcription factors Positive regulation Driving ULK1 mRNA and protein expression in severe hypoxia and ER stress 47
10 STAT1 Transcription factors Negative regulation STAT1 bound a putative regulatory sequence in the ULK1 5′-flanking region, the mutation of which increased ULK1 activator activity 48
11 DAXX Transcription factors Negative regulation Repressing expression of Ulk1 in vivo 49
12 CREB Transcription factors Positive regulation Up-regulation of autophagy genes, including Atg7, ULK1, and Tfeb, by recruiting the coactivator CRTC2 50
13 Nr1d1 Protein Positive regulation Nr1d1 activates ULK1 expression by engaging directly on the distal ROREs to enhance its transcription. 51
14 STRA8 Protein Negative regulation STRA8 binds to the Nr1d1 promoter to inhibit Nr1d1 expression thus to inhibit Nr1d1-regulated ULK1 up-regulation 51
15 p53 Protein Positive regulation Phospho-p53 Ser392 increases Ulk1 expression by binding to ULK1 promoter 52
16 HMGA1 Protein Negative regulation HMGA1 binds ULK1 promoter region and negatively regulates its transcription 53
17 HMGN2 Protein Positive regulation HMGN2 promots the transcription of ULK1 54
18 H3K27ac Protein Negative regulation H3K27ac inhibits the transcription of ULK1 54
19 PRPF8 Spliceosome Positive regulation Regulating Ulk1 mRNA splicing 55
20 PGC-1α Transcription co-activators Positive regulation PGC-1α may regulate ULK1 expression in coordination with ERRα 56
21 DDX3 ATP-dependent RNA helicase Positive regulation DDX3 promotes ULK1 expression post-transcriptionally, which may act by binding to ULK1 mRNA 58
22 miR-372 MicroRNA Negative regulation Decreasing the ULK1 expression by binding sequences in the 3′ UTR of ULK1 61
23 miR-93 MicroRNA Negative regulation Reducing the protein levels of ULK1 under hypoxia condition 62
24 miR-142-5p MicroRNA Negative regulation Inhibiting translation of ULK1 mRNA though binding to the ULK1 3′UTR 63,102
25 miR-20a MicroRNA Negative regulation Decreasing the endogenous ULK1 protein levels by binding sequences in the 3′ UTR of ULK1 64,65
26 miR-106b MicroRNA Negative regulation Decreasing the endogenous ULK1 protein levels by binding sequences in the 3′ UTR of ULK1 65
27 miR-106a MicroRNA Negative regulation Down-regulating ULK1, ATG7, and ATG16L1 proteins 66, 67, 68
28 miR-26a/b MicroRNA Negative regulation Suppressing the post-transcriptional ULK1 expression 69
29 miR-489 MicroRNA Negative regulation Diminishing the expression of ULK1 and LAPTM4B 70
30 miR-1262 MicroRNA Negative regulation Down-regulating of ULK1 expression 71
31 miR-214 MicroRNA Negative regulation Suppressing ULK1 expression through direct binding at a 3′ untranslated region sequence 72
32 miR-155 MicroRNA Negative regulation Down-regulating the expression of ULK1, FoxO3, atg14, atg5, atg3, lc3 and GABARAPL1 73
33 miR-20a-5p MicroRNA Negative regulation Down-regulating the expression of endogenous ULK1 74
34 PVT1 Long non-coding RNA Positive regulation PVT1 positively regulates ULK1 by inhibiting miR-20a-5p 74
35 miR-558 MicroRNA Negative regulation Down-regulating the expression of Ulk1 75
36 MALAT1 Long non-coding RNA Positive regulation MALAT1 positively regulates ULK1 by inhibiting miR-558 75
37 SNHG6 Long non-coding RNA Positive regulation SNHG6 positively regulates ULK1 by inhibiting miR-26a-5p 76
38 miR-26a-5p MicroRNA Negative regulation Down-regulating the expression of ULK1 through directly acting on the 3′UTR of ULK1 76, 77, 78
39 miR-30b-3p MicroRNA Negative regulation Down-regulating the expression of Ulk1 79
40 Gm15834 Long non-coding RNA Positive regulation Gm15834 positively regulates ULK1 by inhibiting miR-30b-3p 79
41 miR-665 MicroRNA Negative regulation miR-665 decelerates the expression of ULK1 by binding to its 3′-UTR 80
42 miR-17 MicroRNA Negative regulation Down-regulating the expression of f PHLPP, ULK1, ATG7 and p62 81
43 PTENP1 Long non-coding RNA Positive regulation PTENP1 positively regulates ULK1 by inhibiting miR-17 and miR-20a 81
44 miR-22 MicroRNA Negative regulation miR-22 directly targets ULK1 to inhibit the expression of the target gene 82
45 miR-34a-5p. MicroRNA Negative regulation Down-regulating the protein expression of ULK1 83
46 circ_0009910 Circular RNA Positive regulation Circ_0009910 activates ULK1-induced autophagy via inhibiting miR-34a-5p 83
47 miR-1275 MicroRNA Negative regulation Decreasing ULK1 and ATG7 expression at RNA and protein level 84
48 circCDYL Circular RNA Positive regulation circCDYL activates ULK1 by inhibiting miR-1275 84
49 miR-514a-3p MicroRNA Negative regulation miR-514a-3p negatively modulates ULK1 expression by direct interaction 85
50 miR-132-5p MicroRNA Positive regulation miR-132-5p positively regulates ULK1 mRNA and protein levels 86
51 miR-25 MicroRNA Negative regulation Knockdown of miR-25 significantly increased ULK1 levels 87
52 miR-106a MicroRNA Posttranscriptional regulation Overexpression of miR-106a resulted in markedly reduced ULK1 expression 88
53 miR-135b-5p MicroRNA Positive regulation Inducing protective autophagy through the MUL1/ULK1 signaling pathway 89
54 miR-378 MicroRNA Positive regulation Activating mTOR/ULK1 pathway and sustaining autophagy 90
55 miR-126 MicroRNA Positive regulation Overexpression of MIR126 induced the AMPK phosphorylation, which in turn activated the ULK1 pathway 91
56 miR-21 MicroRNA Positive regulation miR-21 regulated autophagy activity via AMPK/ULK1 signaling pathway 92
57 miR-99 MicroRNA Positive regulation miR-99 family promoted autophagy through mTOR/ULK1 signaling 93
58 miR-3473b MicroRNA Negative regulation Down-regulating the expression of TREM2, Ulk1 and inhibiting autophagy 94
59 miR-125b MicroRNA Negative regulation Down-regulating the expression of ULK1 96
60 HOTAIRM1 Long non-coding RNA Positive regulation HOTAIRM1 positively regulates ULK1 by inhibiting miR-20a, miR-106b, miR-125b 96
61 NEAT1 Long non-coding RNA Positive regulation NEAT1 knockdown inhibited the protein expression of ULK1 97
62 GAS5 Long non-coding RNA Positive regulation Overexpression of GAS5 upregulated ULK1/2 protein levels 98
63 HOTAIR Long non-coding RNA Positive regulation HOTAIR targets AMPK/mTOR/ULK1 pathways 99
64 PURPL Long non-coding RNA Negative regulation Over-expression of PURPL increases the level of p-ULK1 (Ser757) and inhibits p-ULK1 (Ser555 and Ser317) to repress autophagy 100
65 circTMEM87A Circular RNA Positive regulation TMEM87A elevates ULK1 by inhibiting miR-142-5p 102
Figure 3.

Figure 3

Transcriptional and post-transcriptional regulation of ULK1. (A) Some genes, DNA-binding proteins, transcription co-activators and transcription factors regulate ULK1 expression. (B) Post-transcriptional regulation of ULK1 via RNA helicase, miRNAs, long non-coding RNAs and circRNAs.

3.1.2. Post-transcriptional regulation

Post-transcriptional gene regulation is another key regulatory mechanism of gene expression, mainly including the process of RNA splicing, RNA interference, MicroRNAs, RNA editing, messenger RNA (mRNA) stability and others57. As to ULK1, its post-transcriptional regulation is also diverse. The current research on the post-transcriptional regulation of ULK1 mainly focuses on the factors that regulation of ULK1 mRNA, including RNA helicase, microRNAs (miRNAs), long non-coding RNAs (lncRNAs) and circular RNAs (circRNAs). Recently, researchers found that the post-transcriptional regulation of Atg1 (the homolog of ULK1 in yeast) was of great importance to autophagy control. In yeast, RNA-helicase Ded1 (DDX3 in mammals) was proved to promote the translation initiation of Atg1 by atg1 mRNA binding. Correspondingly, mammal DDX3 could up-regulate ULK1 expression without affect ULK1 mRNA level, indicating this effect was post-transcriptional regulation58.

Importantly, miRNAs are largely contributed to the post-transcriptional regulation of ULK1. miRNAs are a series of endogenous small non-coding RNAs, approximately 17–24 bases in length, that bind to target mRNAs to achieve gene silencing and translational repression59. Uridine at the 5′ end of miRNA, which can partially complement the 3′ untranslated region of mRNA, leading to in translation repression of the mRNA, thus, playing a vital role in post-transcriptional regulation of gene expression60. Numerous studies have unveiled that multiple mi-RNAs, including miR-37261, miR-9362, miR-142-5p63, miR-20a64,65, miR-106b64, miR-106a66, 67, 68, miR-26a/b69, miR-48970, miR-126271, miR-21472, miR-214-3p46, miR-125b72, and miR-15573 inhibit ULK1 gene expression by directly targeting the mRNA of ULK174, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88. In addition, microRNAs could also play an indirect role by regulating ULK1-related pathways. For example, miRNAs, including miR-135b-5p89, miR-37890, miR-12691, miR-2192 and miR-9993 positively regulate ULK1-related pathways, whereas miR-3473b act as a sponge in ULK1-related pathways. Of them, miR-378 initiates autophagy via mTOR/ULK1 signaling pathway and inhibits apoptosis in a cell-autonomous manner90. Importantly, miRNAs are also closely related to human disease. For instance, miR-135b-5p levels are elevated in oxaliplatin-resistant colorectal cancer. Mechanistically, miR-135b-5p induced protective autophagy in cancer cells via activation of MUL1/ULK1 signaling pathway, leading to oxaliplatin resistance. Interestingly, this provided a potential autophagy-based therapeutic strategy for preventing drug resistance89. And miR-3473b, which can indirectly negative regulation of ULK1 by inhibiting TREM2, exerting a regulatory role in autophagy in the etiology and pathogenesis of inflammation in Parkinson's disease (PD)94.

Notably, lncRNAs may contribute a lot to the post-transcriptional regulation of ULK1. LncRNA, as a critical regulatory factor in various cellular processes, is subject to post-transcription monitoring to obtain sufficient stability to maintain the cellular environment and control different physiological functions and plays a critical part in the occurrence, progress, and change of multiple human diseases95. As to ULK1, the effect of lncRNA on ULK1 mainly achieved by regulating different miRNAs, and they all positively regulate ULK1. For instance, PVT1 inhibits miR-20a-5p and thus modulates ULK1 expression74. MALAT1 functioned as decoys sponging miR-558 to up-regulate ULK1 levels75. Moreover, in 5-fluorouracil (5-FU)-resistant CRC cells, SNHG6 facilitates chemoresistance by antagonizing miR-26a-5p and targeting ULK1 to promote autophagy76,77. And in hepatic carcinoma (HCC) cells, PTENP1 could indirectly restrain autophagy by inhibiting the expression of the autophagy gene ULK1 via sponging miR-17, miR-19b, and miR-20a, inducing hepatocyte tumor growth81. Similarly, HOTAIRM1 can antagonize miR-20a, 106b, miR-125b and positively regulates their target ULK196. In a cardiac hypertrophy model, overexpression of miR-30b-3p inhibited autophagy-induced cardiac hypertrophy by binding endogenously to the 3′UTR of ULK1, and lncRNA Gm15834 targeted miR-30b-3p to inhibit this process and promoted cardiomyocyte autophagy, which may be the root cause of aggravating cardiac hypertrophy79. NEAT1 targets miR-34a and promotes the expression of ULK1 to promote autophagy to enhance the sensitivity of CRC to 5-FU97. GAS5, an anti-tumor lncRNA, positively correlates with ULK1/2 and enhances the sensitivity of tumors to chemotherapeutics in an autophagy-independent manner98. In addition, lncRNA can affect autophagy and other physiological functions by acting indirectly on UlLK1 as well. HOTAIR, as a typical example, targets AMPK/mTOR/ULK1 pathways to upregulate autophagy99. PURPL enhances physical interaction with mTOR and ULK1 to promote the level of p-ULK1 (Ser757) mediated by mTOR, thereby inhibiting autophagy to promote cutaneous melanoma100.

Besides, circRNAs are an emerging hotspot for the post-transcriptional regulation of ULK1. circRNAs composed of a covalently closed-loop structure, and characterized by the absence of 5′cap and 3′tail101. Of them, CircCDYL, an autophagy-associated circRNA, promotes the development of breast cancer through miR-1275–ATG7/ULK1 pathway and may serve as a promising predictive molecule of breast cancer84. And in gastric cancer (GC), TMEM87A could elevate ULK1 via sponging miR-142-5p to contribute to cancer cell proliferation and metastasis102. In addition, up-regulation of circ_0009910 was detected in the serum of imatinib-resistant chronic myeloid leukemia patients, and this up-regulation enhanced ULK1-regulated autophagy by targeting miR-34a-5p to promote drug resistance of cancer cells83 (Table 1, Fig. 3B). Therefore, post-transcriptional regulation has a significant impact on the expression and activity of ULK1, and elucidating the effect of post-transcriptional regulation on ULK1 will help to understand the importance of ULK1 in human diseases.

3.2. Post-translational modifications of ULK1 signaling network

A variety of proteins constitute an exquisite PPI network to regulate and be regulated by ULK1, involving multiple different signaling pathways and participating in a variety of life processes. Of note, these regulations are mainly post-translational modifications such as phosphorylation, ubiquitination, methylation, acetylation, glycosylation and sulfhydration (Fig. 4 and Table 2). In this section, we aimed to review the post-translational modifications of ULK1-network in canonical autophagy, non-canonical autophagy, mitophagy and non-autophagy functions.

Figure 4.

Figure 4

Post-translational modifications that regulation of ULK1 signaling network. (A) Post-translational modifications of ULK1 signaling network in canonical autophagy. In the initial stage of autophagy, mTORC1 initiates autophagy by interacting with ULK1 complexes formed by ULK1, ATG13, and FIP200. Then, the PI3K–Beclin1 complex initiates autophagy vesicle nucleation, and ATG12–ATG5–ATG16L1 binds to the vesicles to mediate the formation of pre autophagosomes. In this process, multiple proteins affect ULK1 through phosphorylation, dephosphorylation, acetylation, ubiquitination, and glycosylation, etc. (B) Post-translational modifications of ULK1 signaling network in non-canonical autophagy. (C) Post-translational modifications of ULK1 signaling network in mitophagy, Parkin ubiquitinates polymeric-ub chains, providing more ubiquitin substrates for PINK1 phosphorylation, and the adaptor proteins OPTN and NDP52 promote the recruitment of ULK1 by binding to LC3, initiating the formation of autophagosomes on the mitochondrial surface. (D) Post-translational modifications of ULK1 signaling network in non-autophagy functions under some conditions (e.g., glycolysis, inflammation, stress, cancer, cell death, etc.).

Table 2.

Post-translational modification of ULK1, the ULK1 complex and ULK1 regulated relevant pathways.

Function The substrate protein Modified site Positive regulation/
negative regulation
Modification enzyme Ref.
Phosphorylation ULK1 Ser317, Ser777, Ser555, Thr574, Ser637, Ser467 Activate AMPK 12,103,106
Phosphorylation ULK1 Ser757/758, Ser638, Ser467 Inhibit mTOR 103,104
Phosphorylation mTOR Ser2481 Inhibit ULK1 107
Phosphorylation S6k1 Thr389 Inhibit ULK1 107
Phosphorylation Raptor Ser863, Ser855, Ser859, Ser792 Inhibit ULK1 107
Phosphorylation ULK1 Ser317, Ser555 Activate MAPK15 108,109
Phosphorylation ULK1 Ser405, Ser415 Activate GSK3β 110
Phosphorylation ULK1 Ser423, Ser317/555/777 Inhibit PKCα 111,112
Phosphorylation ULK1 Ser555, Ser757 Inhibit WNK1 113,114
Phosphorylation ULK1 / Activate PP2A 115
Deubiquitination ULK1 / Stabilize USP20 116
Ubiquitination ULK1 / Destabilize USP24 117
Deubiquitination ULK1 / / USP1 118
Ubiquitination ULK1 Lys925, Lys933 Inhibit NEDD4L 119
Ubiquitination ULK1 Lys63 Activate TRAF6 120
Ubiquitination ULK1 / Activate TRAF6-AMBRA1 121
Ubiquitination ULK1 / Activate TRIM32-AMBRA1 122
Deubiquitination ULK1 Lys48 Stabilize STAMBP/AMSH 123
PolyUbiquitination ULK1 K63 Activate NGF 124
Ubiquitination ULK1 K63 Activate TRIM16 125
Phosphorylation TRIM16 Ser116, Ser203 Activate ULK1 125
Dimethylation ULK1 Arginine170 Activate PRMT5 126
Demethylation ULK1 Arginine170 Inhibit KDM5C 126
Acetylation ULK1 Lys162 and Lys606 Activate TIP60 127
Deacetylation ULK1 Lys68 Inhibit HDAC2 128
Glycosylation ULK1 Thr754 Activate OGT 129
Sulfhydration ULK1 Cys951 Activate CSE 130
Phosphorylation ATG14 Ser29 Activate ULK1 131
Phosphorylation Beclin-1 Ser30 Activate ULK1 132
Phosphorylation VPS15 Ser861, Ser865 Activate ULK1 133
Phosphorylation mATG9 Ser14 Activate ULK1 134
Phosphorylation DENND3 Ser554, Ser572 Activate ULK1 135
Phosphorylation PIKFYVE Ser1548 Activate ULK1 136,137
Phosphorylation ULK1 Ser1042/Thr1046 Activate ULK1 139
Ubiquitination ULK1 / Activate KLHL20 139
Phosphorylation SEC23B Ser186 Inhibit ULK1 140
Phosphorylation AMBRA1 / Activate ULK1 141
Phosphorylation ATG4B Ser316 Activate ULK1 142
Phosphorylation FLCN Ser406, Ser537, Ser 542 Activate ULK1 143
Phosphorylation PPP2/PP2A / Activate ULK1 144
Phosphorylation ULK1 Ser317 Activate SQSTM1/p62 145
Phosphorylation ULK1 Ser317 Activate PERK 146
Phosphorylation PERK Ser317 Activate ULK1 146
Dephosphorylation ULK1 Ser637 Activate PPM1D 147
Phosphorylation ULK1 Ser556 Activate DAPK3 148
Phosphorylation ULK1 Ser469, Ser495, Ser533 Inhibit TOPK 149
Phosphorylation ULK1 Ser504, Ser757 Inhibit p38 MAPK 150
Phosphorylation ULK1 D485 cleave Caspase 3 151
Phosphorylation SQSTM1/p62 Ser409 Activate ULK1 152
Phosphorylation ATG16L1 Ser278 Activate ULK1 153
Phosphorylation SEC16A Ser846 Activate ULK1/2 154
Ubiquitination ULK1 / Activate MUL1 155
Phosphorylation ULK1 / / MAPK1/3 156
Phosphorylation ULK1 / Activate NS1 157
Phosphorylation FUNDC1 Ser17 Activate ULK1 158
Phosphorylation Rab9 Ser179 Activate ULK1 160
Phosphorylation FliI Ser64 Inhibit ULK1 161
Phosphorylation ULK1 / Activate TBK1 163
Phosphorylation TBK1 Ser172 Activate ULK1 164
Phosphorylation ULK1 Ser774 Activate AKT 165
Ubiquitination ULK1 Lys48 Inhibit TRAF3 166
Phosphorylation ULK1 Ser555 Activate p38α 167
Phosphorylation ULK1 / Activate CDN 168
Phosphorylation LARS1 Ser391, Ser720 Inhibit ULK1 169,170
Phosphorylation HK Ser124, Ser364 Activate ULK1/2 171
Phosphorylation PFK1 Ser74, Ser762 Activate ULK1/2 171
Phosphorylation ENO1 Ser115, Ser282 Activate ULK1/2 171
Phosphorylation FBP1 Ser63, Ser88 Activate ULK1/2 171
Phosphorylation Dsh Ser239, Ser247, Ser254, Ser266, Ser376, Ser554, Ser555 Inhibit ULK1 173
Phosphorylation Exo70 Ser89 Inhibit ULK1 174
Phosphorylation Cdc37 Ser339 Activate ULK1 175
Phosphorylation Mad1 Ser546 Activate ULK1 176
Phosphorylation STING Ser366 Inhibit ULK1 178
Phosphorylation RIPK1 Ser357 Inhibit ULK1 180

3.2.1. Post-translational modifications of ULK1-network in canonical autophagy

3.2.1.1. Upstream pathways

ULK1, known as the initiator of canonical autophagy, many proteins control the process of autophagy by regulating its activity in various ways (Fig. 4A). Among them, the most well-known is the regulation of ULK1 activity by AMPK and mammalian target of rapamycin (mTOR) via phosphorylation. The cell energy sensor AMPK sustains cell energy balance by sensing the cell energy state and promotes autophagy; on the contrary, mTOR exerts an inhibitory effect in autophagy and is a regulator of cell growth and nutritional signal transduction. Under nutrient deprivation, AMPK could activate ULK1 via directly phosphorylation ULK1 at Ser317, Ser777, Ser555, Thr574, Ser637 and Ser467 to promote autophagy12,103,104. While when nutrient sufficiency, ULK1 could be phosphorylated at Ser757 and Ser638 by mTORC1 directly to disrupt AMPK–ULK1 interaction to impede ULK1-induced autophagy activation further; interestingly, this signal may inactivate ULK1 by phosphorylation at Ser467103, 104, 105. And ULK1 can phosphorylate TOR1-regulated protein (Raptor), which prevents mTOR from binding to substrate Raptor, and thus inhibiting the activation of mTORC1 signaling pathway12,106. Intriguingly, ULK1 could affect S6K1 (substrate of mTORC1) Thr389-specific kinase or phosphatase in Drosophila and mammalian cells107. In addition, ULK1 over-expression increases the phosphorylation of Raptor at Ser863, further critically inhibits phosphorylation of S6K1 and 4e-BP1 mediated by mTORC1, which reduce Raptor's affinity binding to 4E-BP1's substrate yet has little effect on the protein–protein interaction of the mTORC1 component107. Under starvation, mitogen-Activated protein kinase 15 (MAPK15) directly stimulated AMPK-dependent ULK1 activity and affected the phosphorylation of ULK1 to activate ULK1108,109. Besides, glycogen synthase kinase 3 (GSK3) β promotes the phosphorylation of ULK1 at Ser405 and Ser415 in gamma-aminobutyric acid receptor-associated protein (GABARAP)-interacting regions. Phosphorylation of these residues promotes the interaction of ULK1 with MAP1LC3B and gamma-aminobutyric acid receptor-associated protein-like 1 (GABARAPL1), thereby inducing autophagic flux110. Moreover, PKCα phosphorylates ULK1 Ser423 to prevent autolysosome formation, during which ULK1 does not alter activity but reduces autophagosome-lysosome fusion by binding to synapsin 17. Further, phosphorylated ULK1 promotes autophagy-mediated by chaperone promoting degradation of ULK1 by interacting with the heat shock cognate 70 kDa protein111,112. However, WNK lysine deficient protein kinase 1 (WNK1) could inhibit autophagy via inhibiting phosphorylation of ULK1113,114. Through phosphorylation proteomic analysis, combined with relevant experimental data, studies found that ULK1 can directly phosphorylate and promote the regulation of PP2A subunit striatum. Activated PP2A enhances autophagy through positive feedback regulation, promotes protein transport, and degrades proteins115. Beside the above phosphorylation regulation, there are many other proteins that regulate the phosphorylation of ULK1 (Table 2), and contribute to the functions of ULK1.

In addition to phosphorylation regulation, ubiquitination regulation is also critical for ULK1 stability and activity. For instance, ubiquitin-specific protease (USP) 20 could inhibit the degradation of ULK1 by lysosomes through deubiquitylation and stabilization of ULK1, and play a positive role in autophagy initiation116. Ubiquitin specific peptidase 24 (USP24) also inhibits autophagic activity by ubiquitinating ULK1 and affecting its stability. Not surprisingly, when USP24 was knocked out, the expression level of ULK1 was significantly increased and the autophagic flux was increased117. However, USP1, as a key player in the modulation of ULK1 K63-linked deubiquitylation, regulates the Lys63 of ULK1, increasing normative autophagy flux and inhibiting alternative pathways, which lead to lysosomal-mediated degradation of sequestosome-1 (SQSTM1/p62)118. In addition, the E3 ligase NEDD4L specifically ubiquitinates Lys925 and Lys933 sites of ULK1 for degradation by the proteasome to prevent cell death from excessive autophagy119. TNF receptor associated factor 6 (TRAF6) ubiquitin ligase induces autophagy by ubiquitinating, stabilizing, and activating Lys63 of ULK1120. Interestingly, AMBRA1 can also interact with TRAF6, ubiquitinate ULK1 with a Lys63-linked chain, strengthen its stability and regulate its function121. When muscle cells atrophy, the ubiquitin ligase TRIM32 binds to AMBRA1 and ULK1 and activates ULK1 through unanchored K63-linked polyubiquitin chains122. In addition, STAM-binding protein (STAMBP/AMSH) stabilizes ULK1 by eliminating K48-linked ubiquitination, thus positively affecting autophagy flux123. Additionally, nerve growth factor (NGF) could control ULK1 activity via interacting with ULK1 to regulate its polyubiquitination124. Besides, tripartite motif-containing protein 16 (TRIM16) could interact with ULK1 and promoted its K63-linked ubiquitination. Interestingly, ULK1 could in turn regulate TRIM6 activity via phosphorylation it at Ser116 and Ser203125.

Notably, recent studies have indicated that methylation, acetylation, glycosylation and sulfhydration modifications are also important for the regulation of ULK1 activity. Recently, researchers found that the methylation of ULK1 is crucial for autophagy process under hypoxia126. Mechanistically, protein arginine methyltransferase 5 (PRMT5) dimethylation of ULK1 at arginine 170 (R170me2s) could enhance the autophosphorylation of ULK1 at Thr180 to activate ULK1. While lysine demethylase 5C (KDM5C) could remove this modification via its demethylase activity to attenuate the activation of ULK1126. Besides, HIV Tat-interactive protein (TIP60), an acetyltransferase, activates ULK1 to regulate autophagy flux in the face of growth factor deprivation. The Ser86 site of the TIP60 is phosphorylated by GSK3, and TIP60 regulates its activity by acetylation of ULK1 at Lys162 and Lys606. Importantly, GSK3–TIP60–ULK1 pathway connecting two different modifications of phosphorylation and acetylation, which is the ingenious and brilliant part of the regulation of cellular protein modification127. Additionally, histone deacetylase 2 (HDAC2) could regulate ULK1 by deacetylation its K68 site, and this modification contributed to pyroptosis regulated by ULK1–NLRP3 pathway in acute liver failure128. Interestingly, when the glucose level is limited, the Thr754 site of ULK1 can be glycosylated by O-linked N-acetylglucosamine transferase (OGT), which is essential for the binding and phosphorylation of ATG14, thereby stimulating phosphatidylinositol-(3)-phosphate to promote the formation of phagocytes and initiates autophagy129. Recently, sulfhydration modification was also found to regulate ULK1 activity and autophagy by cystathionine gamma-lyase (CSE) sulfhydration of ULK1 at Cys951130.

3.2.1.2. Downstream pathways

Correspondingly, as the initiator of autophagy, ULK1 indisputably regulates many downstream proteins to perform multiple processes of autophagy, through various signaling pathways (Fig. 4A). For instance, ULK1 phosphorylates ATG14 at Ser29 through mTOR pathway, promotes lipid kinase activity of Vps34, and activates autophagy131. Further, ULK1 phosphorylates Ser30 of beclin-1, and activates the PIK3C3 complex containing ATG14, thus promoting the formation of autophagosomes under amino acid starvation, hypoxia, and mTORC1 inhibition132. Another member of the Vps34 complex, VPS15 can also be phosphorylated at Ser861 and Ser865 by ULK1, which enhance its autophagy-promoting ability133. Under the circumstance of nutrient limitation, ULK1 phosphorylate Ser14 sites of mATG9, synergistically promote the interaction between mATG9 and adaptor protein 1/2; induce mATG9 autophagy vesicle transport; further initiating autophagy134. Also, during starvation, ULK1 phosphorylates Ser554 and Ser572 sites of guanine nucleotide exchange factor DENN domain-containing protein 3 (DENND3), which increased its activity on GTPase Rab12. Activated Rab12 binds to LC3 to promote the transport of autophagosomes in starvation-induced autophagy135. Additionally, ULK1 can phosphorylate and activate the lipid kinase PIKFYVE Ser1548 in the face of intracellular glucose starvation, thereby promoting the formation of autophagosomes containing phosphatidylinositol 5-phosphate, thus driving the up-regulation of autophagy136, 137, 138. ULK1 is also a substrate for Cul3-KLHL20 ubiquitin ligase. During starvation, ULK1 autophosphorylation promotes the ubiquitination and proteolysis of the Kelch-like protein Klhl20, contributing to autophagy termination139. What's more, in the case of nutritional deficiencies, ULK1 phosphorylates SEC23B Ser186 to prevent the interaction of SEC23B with the repeat protein FBXW5, thereby stabilizing SEC23B. Phosphorylation and stabilization of SEC23B can enhance its combine with SEC24A and SEC24B, relocate to the ER-Golgi intermediate compartment, and further increase autophagy flux140. Also, in ER, ULK1 phosphorylates AMBRA1 and regulates the dissociation of AMBRA1-DLC1 from the dynein complex, and of note, AMBRA1-DLC1 relocates to the ER when autophagy begins141. In addition, ULK1 directly phosphorylates and inhibits human ATG4B on Ser316 to inhibit LC3 processing. Conversely, the phosphatase PP2A-PP2R3B could eliminate this process. They cooperate to regulate the LC3 process by governing the cellular activity of ATG4B142. Moreover, folliculin (FLCN) is closely related to the cellular pathway and is a novel autophagy element. The absence of FLCN will moderately damage the basic autophagy flux. ULK1 phosphorylates Ser406, Ser537, and Ser542 of the FLCN complex, which interacts with GABARAP, the second component of the autophagy machinery that also regulated by ULK1143. Besides, based on unbiased phosphor-proteomic experiments, it was found that ULK1 phosphorylated the STRN subunit of protein phosphatase 2 (PPP2/PP2A), which in turn supports PPP2 activation through a positive feedback loop and increases protein degradation by promoting autophagy144.

3.2.2. Post-translational modification of ULK1-network in non-canonical autophagy

3.2.2.1. Upstream pathways

ULK1 can also participate in regulating some non-canonical autophagy signaling pathways (Fig. 4B), such as stress, inflammation, toxic reactions, etc. SQSTM1/p62 induces phosphorylation of ULK1 at Ser504 and Ser757, promotes reciprocal regulation of AMPK and ULK1, and induces autophagy to degrade kelch like ECH associated protein 1 (KEAP1) and activate NFE2L2/NRF2, which rescues lipotoxicity in mouse liver cells145. Also, PKR-like ER kinase (PERK) can directly phosphorylate ULK1 Ser317 and subsequently induce autophagic KEAP1 degradation in response to lipotoxicity146. Notably, protein phosphatase 1D magnesium-dependent delta isoform (PPM1D) does not participate in the autophagy process induced by starvation, it regulates ULK1 activity by dephosphorylating ULK1 Ser637 in a p53-dependent manner and formats ULK1 puncta to activate autophagy147. Further, multiple proteins exert anti-tumor effects through phosphorylation of ULK1. Death-associated protein kinase 3 (DAPK3) can directly promote the formation of ULK1 complex by phosphorylation at Ser556, which contribute to the anticancer role of DAPK3 in GC148. T-LAK cell-derived protein kinase (TOPK) could also inhibit the activity and stability of ULK1 by phosphorylating ULK1 (Ser469, Ser495, and Ser533), thereby reducing autophagic activity and decreasing the sensitivity to temozolomide of glioma149. What's more, p38/MAPK triggers the phosphorylation of ULK1 at Ser504 and Ser757, preventing its binding to downstream effector protein ATG13, inhibiting autophagy in microglia, and promoting microglial inflammation150. Intriguingly, apoptotic executive protein caspase-3 could cleave ULK1 on D485 site thus to control ULK1 activity in leukemogenesis151.

3.2.2.2. Downstream pathways

Proteotoxic stress promotes phosphorylation of SQSTM1/p62 at the ubiquitin-association (UBA) domain of toxins, regulating its binding to ubiquitinated proteins. Mechanistically, when nutrients are abundant, ULK1 phosphorylates SQSTM1/p62 at UBA Ser409, thereby destroying the stability of the UBA dimer interface as well as increasing SQSTM1/p62 binding to ubiquitin152. ULK1 also activates autophagy by phosphorylation of downstream proteins to participate in the regulation of some stress mechanisms. Excepted for lapidating LC3B and driving the formation of autophagosomes, ATG16L1 is also a direct ULK1 target and regulated by ULK1 phosphorylation at Ser278 during stress153. In addition, based on proteomic methods combined with experimental methods, it was found that ULK1 phosphorylated SEC16A at Ser846, promoting the assembling COPII components in ERES, and regulating the transport of ER to the Golgi apparatus154.

3.2.3. Post-translational modification of ULK1-network in mitophagy

3.2.3.1. Upstream pathways

Notably, ULK1 also has a very important role in the selective autophagy of cells, of which mitophagy is the classical type (Fig. 4C). ULK1 often acts as an upstream regulator to regulate mitophagy, but there are still some proteins that can affect the process of mitophagy by affecting ULK1. For instance, in selenite-induced mitophagy, ULK1 acts as a substrate for mitochondrial E3 ubiquitin ligase 1 (MUL1). MUL1 ubiquitinates ULK1 and promotes ULK1 degradation to regulate selenite-induced mitophagy155. Notably, functions of ULK1 in mitophagy also affect the disease's progression of cancer, inflammation, and viral infections. For example, abnormal levels of ULK1 expression are frequently observed in breast cancer, which may be due to the degradation of Lys48-linked ULK1 ubiquitination by MAPK1/3 kinases in breast cancer. Not surprisingly, disrupting the MAPK1/3 kinases to increase the phosphorylation of ULK1 and inducing mitophagy, which works well in the treatment of bone metastases with low ULK1 levels156. Nonstructural protein 1 (NS1) protein could also trigger the phosphorylation of ULK1 and enhances the mitochondrial expression of BCL2-interacting protein 3, which were vital for influenza virus mediated mitophagy157.

3.2.3.2. Downstream pathways

To regulate mitophagy, ULK1 could increase the phosphorylation of FUNDC1 at Ser17, thereby enhancing the binding of FUNDC1 to LC3158. Furthermore, the interaction of Sestrin-2 with ULK1 promotes ULK1-mediated phosphorylation of Beclin1 at Ser14, promoting Parkin translocation to the mitochondrial membrane and activating mitophagy159. Further studies revealed that ULK1 phosphorylates receptor-interacting Ser/Thr protein kinase 1 Rab9 at Ser179, thereby recruiting Rab9-related trans-Golgi membranes to impaired mitochondria. This is a pathway mediated by mitophagy that protects the heart from ischemia by degrading damaged mitochondria160. Additionally, Flightless-I (FliI), a p62 interacting protein, was impeded by ULK1 phosphorylation at Ser64 while promoted by Akt-mediated phosphorylation at Ser436. Of note, its inhibition of selective autophagy contributes to breast cancer progression161. Further, Sestrin-2, a stress-inducible protein, interacts with ULK1 to promote phosphorylation of p62 at Ser403 and activate p62-mediated mitophagy162. Intriguingly, autophagy receptor protein NDP52 could interact with the ULK1 complex through FIP200, and this was proved to participate in Parkin-regulated mitophagy. In this process, TANK-binding kinase 1 (TBK1) enhanced this interaction by phosphorylation of NDP52 and recruited the ULK1 complex to ubiquitinate cargo, inducing mitophagy163,164. More interestingly, when ULK1 was activated by AMPKα, activate ULK1 could directly phosphorylation of TBK1 at Ser172, which in turn enhanced the activity of TBK1164.

3.2.4. Post-translational modification of ULK1-network in non-autophagy functions

3.2.4.1. Upstream pathways

ULK1 can participate in regulating some non-autophagy-related signaling pathways (Fig. 4D). For instance, the high-intensity of Akt is located at Ser774 site of ULK1, and the phosphorylation of this site is activated by insulin. Mutation of ULK1 at the Ser774 Akt consensus site inhibits insulin-dependent ULK1 phosphorylation165. What's more, after complexing with TRAF2 and cIAP1, TRAF3 ubiquitinates K48-linked ULK1 to degrade ULK1, a process that regulates macrophage inflammation and pyroptosis166. Besides, the stress kinase p38α can phosphorylate ULK1 at Ser555, and this is important for the cell's destiny, step into senescence or apoptosis167. In addition, ULK1 was found to be highly phosphorylated by key cell cycle regulators cyclin-dependent kinase 1/cyclin B (CDN) throughout the cell cycle168.

3.2.4.2. Downstream pathways

Under glucose starvation, AMPK phosphorylated and activated ULK1, which further phosphorylation of LARS1 at Ser391 and Ser720, thereby reducing the effect on the downstream protein RagD and inhibiting ATP leucine-involved protein translation169,170. Likewise, ULK1 is an excellent energy signaling hub. In the absence of nutrients required by cells, ULK1 directly phosphorylation of critical glycolytic enzymes such as hexokinase (HK), phosphofructokinase 1 (PFK1), enolase enzyme 1 (ENO1) and gluconeogenic enzyme-1,6-bisphosphatase (FBP1), which further activated the glycolytic pathway to maintain glucose metabolism171. Of note, Disheveled (Dsh) plays a central role in the Wnt/β-catenin pathway, which is a key pathway that determines the differentiation fate of cells during development172. Interestingly, ULK1 can phosphorylate Dsh at multiple sites, such as Ser239, Ser247, and Ser254 in PDZ-DEP, as well as Ser239, Ser247, Ser254, Ser266, Ser376, and Ser554173. Moreover, ULK1 is of great importance in cancer-related signaling pathways. For instance, the exocyst subunit Exo70 is phosphorylated by ULK1 at the Ser89 site and thus inhibits cancer cell metastasis174. Additionally, ULK1 mediates phosphorylation of co-chaperone cell division cycle protein 37 (CDC37) at Ser339, which impairs CDC37-coordinated HSP90 function and contributes to degrading HSP90–CDC37 client kinases, thereby increasing the lethality of HSP90 inhibitors to tumor cells175. Further, Mad1, the spindle assembly checkpoint protein, is phosphorylated by ULK1 at Ser546 for recruitment of Mad1 to kinetochores to maintain chromosome fidelity during cancer cell mitosis176. Importantly, ULK1 also involved in the regulation of several inflammatory and stress related pathways. VCP/p97 can be phosphorylated at Ser13, Ser282, and Thr761 by ULK1/2 localized to stress granules to treat VCP mutation-induced IBM-like disease by promoting VCP activity and disaggregating stress granules177. Circulating dinucleotide could also promote the function of the stimulator of interferon genes (STING), which transports TBK1 to the endosome/lysosome and motivates interferon regulatory factor 3 as well as nuclear factor 1 progenitor κB (NF-κB). ULK1 was proved to phosphorylated STING at Ser366, which subsequently prevented the persistent transcription of innate immune genes178. Strikingly, ULK1 is also involved in the regulation of the necroptosis pathway. The receptor-interacting kinase 1 (RIPK1) is a key mediator of TNF-induced signaling, in the process of cell life, RIPK1 acts as the gatekeeper of cell life and death. Particularly, it can decide the destiny of the cell, and the way of cell death179. Intriguingly, ULK1 could phosphorylation of RIPK1 at Ser357, which impeded RIPK1-mediated cell death180.

3.3. Other regulation of ULK1

Interestingly, in addition to transcriptional regulation, post-transcriptional regulation, and post-translational regulation, there are many other regulators that can affect the function and activity of ULK1 (Table 3). Metal copper (Cu), which always acts as the static cofactor within enzyme active sites, was found to contribute to the activation of ULK1/2 through directly interaction. And this interaction was associated with ULK1-mediated autophagosome formation and the autophagy progression181. Intriguingly, researchers found GABARAP and GABARAPL1 could co-regulate ULK1 activity with LC3B/C. Mechanistically, GABARAP and GABARAPL1 bind to ULK1 and ATG13 and enhance ULK1 activity to promote autophagy. On the contrary, LC3B and LC3C decrease the activity of ULK1 through inhibiting GABARAPL1 lipidation182. Moreover, some regulators can affect the activity of ULK1 by regulating the localization of ULK1. For instance, the protein TFG encoded by Trk fusion gene (TFG) plays a pivotal role in autophagy initiation by interacting with ULK1, increasing the LC3C–ULK1 binding183. S100A10, a member of the S100 protein family, was found to directly interact with ULK1 and further stimulate the localization of ULK1 to ER-mitochondria contact sites184. Interestingly, hypoxia can stimulate ULK1 to translocate into nucleus, further regulating YAP-driven glycolysis in pancreatic ductal adenocarcinoma (PDAC)185. Additionally, some regulators enhance the interaction of ULK1 with its partner to affect ULK1 and autophagy. C9orf72, whose mutations are the most common genetic causes for amyotrophic lateral sclerosis (ALS) and frontotemporal dementia (FTD), was proved to function as a component of the ULK1 complex by directly interacting with ATG13 in the ULK1 complex186. The adaptor NDP52 triggers membrane recruitment of the ULK1 complex to initiate mitophagy187. And Rab GTPase SMCR8 could negative regulation of ULK1 kinase activity via binding to ULK1 complex components188. GTPases RAB2 can also interact with the ULK1 complex, thereby promoting the recruitment of the ULK1 complex to form autophagosomes189. Interestingly, NPM1 mutant NPM1-mA could interact with ULK1 protein and positively regulate ULK1 protein levels and contribute to leukemic cell survival190. IRGM, encoded by a human gene for inflammatory disease risk, plays anti-inflammatory roles by promoting the interaction and co-assembly of ULK1 and Beclin 1 and the formation of the ULK1 complex191. What's more, the Golgi protein SCOC can interact with the ULK1-binding protein fiber bundles and the elongation protein zeta 1 (FEZ1). Subsequently, SCOC, ultraviolet radiation-related genes, and FEZ1 form a starvation-sensitive trimeric complex that regulates the activity of ULK1. Of note, during glucose starvation, SCOC promotes autophagy by stabilizing the ULK1–FEZ1–SCOC complex192. Atlastin 2 and 3 (ATL2/3), as ER-localized transmembrane proteins, could recruit and stabilize ULK1 and ATG101, thereby transporting to ER to form autophagosomes designated by FIP200–ATG13193. Besides, TRIM5α is an antiretroviral protein that has been shown to interact with upstream proteins required to initiate autophagy, such as ULK1, to protect cells from infection. Mechanistically, TRIM5α interacts with ULK1 and affected cytoplasmic distribution of active ULK1 (p-ULK1Ser-317)194,195. Intriguingly, cytosolic Malate dehydrogenase 1 (MDH1) has been shown to indirectly regulate ULK1 via regulation of its proteasomal degradation, and when autophagy was activated, the activity of MDH1 could be in turn activated196. And a member of the claudin family protein CLDN1, nucleoporin TPR and metastasis-associated colon cancer-1 (MACC1) were proved to indirectly regulate ULK1 through AMPK and mTOR pathway197, 198, 199. Moreover, tRNA m7G methyltransferase complex proteins METTL1 was found to facilitate tumorigenesis through RPTOR–ULK1 axis. Mechanistically, RPTOR, which is a regulatory associated protein of mTOR complex 1, regulated ULK1 activity through mTOR phosphorylation200 (Fig. 5).

Table 3.

Other factors regulation of ULK1.

No. Name Classification Positive regulation/
Negative regulation
Mechanism Ref.
1 Copper Metal ion Positive regulation Enhancing ULK1 kinase activity by directly interaction 181
2 GABARAP/GABARAPL1 Protein Positive regulation GABARAP/GABARAPL1 positively regulate starvation-induced ULK1 activation and are important for maintaining the expression or stability of the ULK1 complex proteins 182
3 LC3B/C Protein Negative regulation LC3B/C might negatively regulate ULK1 by reducing the expression or stability of GABARAPs 182
4 TFG Protein / Binding LC3C to regulate ULK1 localization and autophagosome formation 183
5 S100A10 Protein Positive regulation S100A10 regulates ULK1 translocation to ER-mitochondria contact sites in response to IFN-stimulation 184
6 Hypoxia Cellular microenvironment Positive regulation Hypoxia stimulates ULK1 to translocate into nucleus 185
7 c9orf72 Protein Positive regulation C9orf72 regulates ULK1 expression via regulation of the ULK1 complex 186
8 NDP52 Protein Positive regulation NDP52 recruits the ULK1 complex and triggers membrane recruitment 187
9 SMCR8 GTPases Negative regulation SMCR8 regulates ULK1 kinase activity, ULK1 gene and protein expression 188
10 RAB2 GTPases Positive regulation RAB2 facilitates the recruitment of ULK1 complex and modulates ULK1 kinase activity 189
11 NPM1-mA Protein Positive regulation NPM1-mA promoted TRAF6-dependent K63 ubiquitination and further maintained ULK1 stability and kinase activity via miR-146a 190
12 IRGM Protein Positive regulation IRGM interacts with ULK1 and Beclin 1 and promotes their co-assembly thus governing the formation of autophagy 191
13 SCOC Protein Positive regulation Activating ULK1 via stabilizing ULK 192
14 ATL2/3 Protein Stabilize Stabilizing the ULK1 complex 193
15 TRIM5α Protein Positive regulation TRIM5a interacts with ULK1 and affected cytoplasmic distribution of p-ULK1 (Ser317) 195
16 MDH1 Protein Positive regulation MDH1 may regulate ULK1 proteasomal degradation 196
17 CLDN1 Protein Positive regulation CLDN1 increases ULK1 expression through AMPK/STAT1 signaling pathway 197
18 TPR Protein Negative regulation TPR regulates ULK1 activity via mTOR 198
19 MACC1 Protein Positive regulation MACC1 regulates ULK1 via AMPK 199
20 METTL1 Protein Negative regulation METTL1 regulates ULK1 via RPTOR, a regulatory associated protein of mTOR complex 1 200

Figure 5.

Figure 5

Other factors that regulation of ULK1-related PPI network. (A) Representative factors regulate ULK1 activity through directly interaction. (B) Representative factors regulate ULK1 activity through regulation of the ULK1 complex. (C) Representative factors regulate ULK1 activity through influence the subcellular localization of ULK1. (D) Representative factors regulate ULK1 activity through regulation of AMPK/mTOR. (E) Representative factors regulate ULK1 activity in some human disease.

In short, the transcriptional regulation, post-transcriptional regulation, post-translational regulation, and other regulatory methods of ULK1 constitute a complex and exquisite cellular program with ULK1 as the core. Through the in-depth exploration of these regulation, we can understand the important role of ULK1 and its precise molecular mechanisms.

4. Autophagic and non-autophagic functions of ULK1 in human diseases

Since the multiple biological functions of ULK1, and ULK1-related signaling pathways have extensive effects on autophagy and other critical cellular processes, its dysregulation can lead to malignancies, neurodegenerative diseases, obesity, heart disease, infections, and other diseases by altering normal human physiology. In this section, we will detail how ULK1 affects the progression of various diseases through its autophagic and non-autophagic functions (Fig. 6).

Figure 6.

Figure 6

The autophagic and non-autophagic functions of ULK1 in cancer, neurodegenerative disease, immune-related diseases, cardiovascular diseases, and other human diseases. The PDB code of ULK1 structure is 4WNO16.

4.1. Cancers

4.1.1. Autophagic function

Autophagy shows a double-edged sword role in different types, stages, and genetic contexts of cancers. In the early phases of tumorigenesis, activation of autophagy might prevent cancer development201. On the contrary, in advanced cancer, both autophagy enhancement and autophagy inhibition have been proposed as therapeutic strategies201. The potential role of autophagy in cancer is of extreme complexity and is always content-dependent, which deserves further discussion202.

On the one hand, autophagy controls the stability of genes and participates in the occurrence, development, treatment, and prognosis of tumors, thereby inhibiting tumors203. Particularly, in tumor tissues with reduced ULK1 expression, promoting autophagy-related cell death by activating ULK1 may be a prospective strategy. Of note, breast cancer is such a typical example, as the expression of ULK1 is remarkably down-regulated, and the baseline of autophagy is reduced as well, indicating that ULK1 might be a promising prognostic biomarker in breast cancer204. The Cancer Genome Atlas (TCGA) and Tissue Microarray (TMA) also exhibit significantly reduced ULK1 expression in breast cancer tissues, especially triple-negative breast cancer (TNBC)15. Mechanistically, the AMPK–ULK1–autophagy pathway correlates with tumor glycolysis orchestrates and tumor immunosuppression maintaining, which are associated with poor human TNBC outcomes205. In addition, ULK1-regulated mitophagy was confirmed to involved in the therapeutic response to targeted drug in breast cancer. For instance, the MAP2K/MEK inhibitor trametinib can up-regulated ULK1 to stimulate mitophagy, which contribute to the reduction of bone metastasis156. Therefore, activation of ULK1 is likely to be a brilliant therapeutic strategy for breast cancer.

On the other hand, autophagy can maintain the protection, survival, and defense mechanisms of cancer cells by keeping mitochondria functional, reducing DNA damage, enhancing cancer cell survival and anti-stress ability, etc., thereby maintaining tumor survival, ultimately promoting tumorigenesis, and causing resistance to therapeutic agents203. Correspondingly, over-expression of ULK1 is negatively correlated with various cancers, including colon cancer, pancreatic ductal adenocarcinoma, liver cancer, prostate cancer (PC), chronic myelomonocytic leukemia, and glioblastoma. Studies have demonstrated that in colon cancer cells, after silencing eukaryotic elongation factor-2 kinase, activating autophagy promotes colon cancer cells survival via AMPK–ULK1 pathway, indicating that the down-regulation of ULK1 might constitute a new treatment for human colon cancer206. Indeed, drugs, such as Cordyceps Sinensis polysaccharide207 and curcumin combined with 5-FU208, have been proved to suppress colon cancer cell proliferation by inhibiting autophagy and apoptosis via impairing AMPK/ULK1 pathway while blocked autophagy flux and lysosome formation. And down-regulating ULK1 could repress tumorigenic activities of pancreatic cancer cells by regulating autophagy-mediated mitochondrial metabolism209. In the hypoxic microenvironment of PDAC, ULK1-triggered autophagy and glycolysis act as a niche for malignant growth and drug resistance by stabilizing previously unrecognized crosstalk between nuclear YAPs during PDAC development185. Notably, PVT1/miR-20a-5p/ULK1/autophagy pathway74, MDH1–ULK1 pathway196, and LKB1–AMPK–ULK1 signaling axis210, may be novel targets for developing therapeutic strategies for PDA by leading to autophagy induction. Besides, based on immunohistochemical analysis of TMAs, ULK1 combination with LC3B is the main prognostic factor for HCC patients211, and ULK1-mediate autophagy was activated when HCC cells faced nutrient deprivation212. In addition, alterations in mitochondrial DNA impairs ROS–AMPK–ULK1 signaling pathways, activating autophagy in HCC cells213. Here, XST-14, a ULK1 inhibitor discovered by molecular docking, can block autophagy, induce apoptosis, and inhibit the development of HCC cells214. Additionally, androgen receptor regulates ULK1 expression in PC patients. Experiments have shown that knocking down ULK1 inhibits androgen-mediated autophagy and PC cell proliferation. Moreover, high levels of ULK1 protein in primary PC are positively related to increased biochemical recurrence, both of which highlight the potential of targeting ULK1-mediated autophagy in the treatment of advanced PC215. Besides, ULK1-mediated pro-survival autophagy can inhibit leukemia cell death in FLT3-ITD mutant leukemia, thus ULK1 may serve as a viable drug target for the treatment of FLT3-ITD mutant leukemia216. Intriguingly, the mRNA expression levels of ULK1 in acute myeloid leukemia (AML) samples were analyzed based on TCGA database, and the level of ULK1 is elevated in AML, which was also associated with poor prognosis in AML217. Moreover, in chronic myeloid leukemia (CML), inhibition of ULK1 could sensitize leukemic stem cells to tyrosine kinase inhibitor (TKI)218. Similarly, P2RY6-mediated autophagic activation of CAMKK2/PRKAA1/ULK1 restores human monocyte differentiation, which is exciting for chronic myelomonocytic leukemia therapy219. In addition, autophagy is also crucial for glioma initiation and growth, destruction of ULK1 by RNAi could inhibition of autophagy and significantly reduced glioblastoma development, indicating that ULK1-mediated autophagy contributes to the survival of tumor220. Notably, ULK1 levels (mRNA and protein) were remarkably increased in GC tissues. The high expression of ULK1 in GC was proved to related to the patient's classification and cancer recurrence. Thus, the overexpression of ULK1 in GC decreases the survival of GC cells and increases the recurrence of cancer. And ULK1 may function as a prognostic biomarker for GC221. In addition, the TCGA data of clear renal cell carcinoma (ccRCC) shows that the mRNA level of ULK1 was much higher in ccRCC tissues than normal kidney tissues. Inhibition of ULK1 by shRNA or ULK1 inhibitors induces cell apoptosis in ccRCC cells, which leading to anticancer efficacy in vivo222. Importantly, AMPK–ULK1–LC3B axis contributes a lot to autophagy inhibition induced anti-cancer effects176. Of note, with the extension of the treatment cycle, drug resistance in cancer cells is generally inevitable223, and ULK1-regulated autophagy also plays a crucial role in drug resistance. For instance, drug resistance of the bromodomain and extraterminal domain inhibitor JQ1 in AML may result from the induction of protective autophagy in leukemia stem cells by activating the AMPK/ULK1 pathway224. In addition, ULK1-mediated autophagy is one of the causes of drug resistance in TKI treatment of CML, which mediated by maintaining energy and redox balance. Therefore, inhibiting ULK1-induced stress differentiation of leukemia stem cells is a practical approach to enhance the therapeutic effect of TKI on CML218,225. Moreover, imatinib resistance in the treatment of CML may be caused by grancalcin activation of ULK1-mediated autophagy120. Besides, osteosarcoma is prone to drug resistance after chemotherapy, attributed to the autophagy induction by enhancing the interaction of high mobility group box 1 protein with Beclin 1 mediated by the ULK1 complex226.

4.1.2. Non-autophagic function

Similarly, the non-autophagy related functions of ULK1 also play a role in cancer. As mentioned in the preceding text, increased expression of ULK1 was negatively correlated with breast cancer metastasis by activating autophagy. Additionally, ULK1 also phosphorylates ERK1/2, which further phosphorylates Exo70, thereby inhibiting the invasion of human breast cancer cells, achieved from a non-autophagic route174. In addition, UlLK1 expression was found to correlate positively with the degree of aggressiveness of the tumor. The spindle checkpoint protein Mad1 is phosphorylated by ULK1, and then p-Mad1 may recruit kinetochores to promote proper mitotic progression, aligns, and segregates faithful chromosomes. Furthermore, evidence shows that ULK1 in cancer enhances the chromosomal stability of paclitaxel and reduces cytotoxicity, thereby promoting tumor cell growth176.

In short, ULK1-mediated autophagy plays different roles in different tumor types and different conditions, acting as the criminal or savior. In general, ULK1-mediated autophagy suppresses cancer in the early stages of tumors while maintaining tumor growth in the late stage of the tumors. Besides, in a minority of advanced cancers, ULK1 also mediates metastasis-related signaling pathways to inhibit cancer invasion, which is a role of ULK1 beyond autophagy. Therefore, based on the complex role of ULK1 in cancer, it is necessary to further clarify the exact role of ULK1 in different situations and carry out corresponding interventions, so that the regulation of ULK1 may play a correct and brilliant role in cancer.

4.2. Neurodegenerative diseases

4.2.1. Autophagic function

Autophagy is closely related to a variety of neurodegenerative disorders, including Alzheimer's disease (AD), PD, Huntington's disease (HD), ALS, frontotemporal lobar degeneration, etc.227. The currently recognized pathogenesis of AD is mainly β-amyloid (Aβ) deposition and neurofibrillary tangles, followed by impaired blood–brain barrier function and neuroinflammation, which may be related to abnormal autophagy. Activation of autophagy can effectively reduce AD-related pathology. Of note, studies found that through inhibition of PI3K/Akt/mTOR/ULK1 pathway and NQO1/mTOR/ULK1 pathway, as well as activation of AMPK/ULK1 pathway, Aβ load and plaques in the brain of AD mice can be reduced, and neuroinflammation and inflammatory bodies can be suppressed, thus improving learning and memory impairment228,229. In addition, impaired steps in autophagy may contribute to PD, and peripheral blood mononuclear cells from PD patients exhibit down-regulated autophagy genes including ULK1 and elevated levels of α-synuclein, suggesting that activation of autophagy may have potential therapeutic effects in PD patients with low basal autophagic activity230,231. Intriguingly, the ULK1 activator BL-918 indeed shown considerable therapeutic potential in cell and animal models of PD232. Additionally, HD is a disease mainly caused by mutations in HTT (huntingtin protein) encoding polyglutamine (polyQ)233. Evidence suggests that mTOR-mediated ULK1 phosphorylation of ATG14 promotes Vps34 lipid kinase activity to clear polyQ mutants in a mouse HD model131. Furthermore, ULK1 phosphorylates SQSTM1/p62 in the UBA domain, and next involves in the clearance of ubiquitinated proteins or polyQ-Htt152. Moreover, multiple studies have demonstrated that C9ORF72 deficiency or mutation will cause FTD and ALS disease, which is also related to ULK1. Mechanistically, C9orf72 can synergize with Rab1a and ULK1 to increase autophagic flux and reduce the accumulation of p62-positive puncta, which may be a relevant pathology in ALS/FTD234. Besides, C9orf72 protein can also form complexes with SMCR8, WD repeat domain 41 (WDR41), and ATG101 to regulate autophagic flux and restore neuronal morphology by regulating ULK1 to regulate autophagy186,235. Additionally, ULK1 also exerts a pathogenic effect in the course of neurodegenerative disease. Chorea-acanthocytosis, an inherited neurodegenerative disease, is characterized by circulating acanthocytes lacking choline. Recent studies have reported that the level of ULK1 and ATG7 is increased in the cytoplasm of chorea-acanthocytosis erythrocytes because HSP forms an HSP90-70 macromolecular complex with ULK1 that sequesters toxic Lyn and hinders its transport, delaying mitochondrial and lysosome clearance236.

4.2.2. Non-autophagic function

Importantly, beyond the autophagy induction role of ULK1, it can also affect neurodegenerative diseases through other ways. ULK1 was found to regulate the process of neuronal development, and its loss of function would contribute to neurodegeneration. Notably, studies have found that ULK1 play a positive role in neuroprotection and regeneration in the central nervous system237. The expression of Ulk1 is up-regulated in the hypothalamus after traumatic brain injury in mice, and knockout of ULK1 can improve the survival of hypothalamic neurons, reduce neuroinflammation and neuronal apoptosis, and reduce glial cell proliferation by inhibiting p38/JNK pathway238,239. Porcine hemagglutinating encephalomyelitis virus (PHEV) is a neurotropic coronavirus that induces neurodegenerative diseases by invading the central nervous system of suckling piglets. A recent study demonstrated that ULK1 functional inhibition was involved in the regulation of PHEV-induced neuronal degeneration, manifested by hypoplastic axonal extension, increased dendritic branching, unstable dendritic spine formation, etc. PHEV infection significantly downregulated the phosphorylation level of the NGF high-affinity receptor TrkA and inhibited ULK1-mediated internalization of NGF, and induced neuronal damage. The up-regulation of mir-142a-5p caused by PHEV infection also negatively regulates NGF signal transduction by inhibiting the expression of ULK1, causing axonal transport disorder and inducing neurodegenerative diseases63,240. It was also found that intracellular protein imbalance can lead to various neurodegenerative diseases, and Ulk1 knockout mice exhibit obvious neuronal degeneration. Mechanistically, ULK1 can regulate ER-to-Golgi cargo transport through phosphorylation of SEC16a to maintain neuronal homeostasis154. In addition, schizophrenia is a grievous, highly inherited mental disease. Based on exome sequence data, exome sequencing of patients with schizophrenia revealed four nonsense mutations in the ULK1 gene, pending independent validation by experiments241. ULK1 also regulates repositioning of the axonal initial segment, which affects potential action generation and axonal polarity242.

Therefore, ULK1 plays a protective role in most neurodegenerative diseases. Mechanistically, ULK1-mediated activation of autophagy can effectively reduce most pathologies associated with neurodegenerative diseases by clearing toxic proteins and reducing toxic proteins neuroinflammation and neuronal apoptosis to lessen brain damage and protect dopamine neurons. However, under certain conditions, ULK1 can play a pathogenic role by blocking the transport of toxic substances.

4.3. Immune-related diseases

4.3.1. Autophagic function

ULK1-mediated autophagy is essential in regulating immune responses, especially against virus infection. Naturally, autophagy acts as a mighty host defense mechanism that inhibits HIV replication. After HIV infection in Jurkat cells and CD4+ T cells, autophagy is induced by increasing the transcription of ULK1 to defend against HIV invasion243. Invading microbial pathogens also can be selectively eliminated by xenophagy. Of note, TBK1 and NDP52 can recruit the ULK1 complex to cytoplasmic bacteria to initiate xenophagy and resist microbial pathogens’ invasion244. Likewise, autophagy can limit the replication of herpes simplex virus 1 in host cells. Unfortunately, to counteract organism defense, HSV-1 Us3 Ser/Thr kinase cooperates with ICP34.5 to regulate the phosphorylation of ULK1 in virus-infected cells to inhibit autophagy and promote HSV replication245. Salmonella enterica could cause severe gastroenteritis, and autophagy-mediated by Atg5 and ULK1 in host cells can resist to the invasion of the virus. However, Salmonella could still withstand the immune response by protecting itself from autophagy. Mechanistically, Salmonella virulence factors SseF and SseG can inhibit Rab1A, thereby reducing ULK1 recruitment and activation, impairing autophagy initiation246,247. Brucella can utilize the Brucella vacuole (BCV) to send it into the ER of the host cell for replication to evade clearance by intracellular autophagy. BCV can be transformed into an autophagy-characterized compartment (aBCV), a process that requires activation of autophagy-initiating proteins such as ULK1, thereby forming the intracellular circulation of Brucella and facilitating subsequent infection of host cells248.

Notably, in some cases viruses or pathogens can take advantage of autophagy-related proteins or products of host cells to assist their replication. For instance, human immune-related GTPase M (IRGM) promotes the co-localization of Golgi vesicles with replicating hepatitis C virus and induces autophagy by promoting ULK1 phosphorylation through IRGM, utilizing cell membrane remodeling, thereby promoting hepatitis C virus replication249. Cryptococcus neoformans, a fungal pathogen that could induce the phosphorylation of LKB1 in host cell, promoted the phosphorylation of autophagy-initiating proteins, including ULK1, which play a role in C. neoformans internalization, trafficking, and replication250. The miR-99 family activates autophagy via IGF-1R-mediated mTOR/ULK1 pathway to promote the replication of hepatitis B virus (HBV)93. Activated AMPK/mTOR–ULK1–autophagy axis promotes HBV replication at low glucose concentrations and predisposes hepatocytes to HBV infection251. In lung inflammation and even acute lung injury under influenza A virus infection, lacking of HIF-1α decreases glycolysis levels and promotes autophagy through the AMPKα/ULK1 pathway to promote IAV copy252. Moreover, under infectious conditions, ULK1 kinase, can directly phosphorylate ATG16L1, and phosphorylated ATG16L1 localizes to the site of internalizing bacteria and promotes antimicrobial autophagy, a process implicated in Crohn's disease, and the mechanism remains to be investigated153.

4.3.2. Non-autophagic function

Interferons are secreted cytokines to defend against viral infections, and ULK1 could regulate type I interferon (IFN)-dependent immunity253. Mechanistically, ULK1 is phosphorylated and activated through the involvement of type I IFNR at ser757, which is dependent on Akt activity, and then activated ULK1 regulates the expression level of IFN-stimulated genes by activating the p38α MAPK pathway, thereby regulating interferon signaling253,254. IFN-γ receptor stimulation could also activate ULK1, and the interaction between ULK1 and the mixed lineage kinase 3 was necessary for mixed lineage kinase 3 phosphorylation and downstream activation of the kinase ERK5. ULK1 promotes the transcription of key antiviral IFN-stimulated genes and is essential for IFN-γ-dependent antiviral effects, which might be a vital element of the antiviral response255. In addition, STING can detect abnormal DNA in the nucleus and directly transcribe and activate type I IFN and NF-κB. Importantly, when persistent DNA damage occurs, ULK1 could inhibit STING activity and reduce the expression of type I IFN by phosphorylating STING at Ser366178. Additionally, ULK1/2 inhibits the formation of SINT puncta and then phosphorylates TBK1, which affects viral infections or protein aggregation diseases256. Besides, the TFG forms a complex with the E3 ligase, TRAF3, and participates in the immune response by destroying the ULK1–TRAF3 interaction to stabilize ULK1 and protect macrophages from pyroptosis257.

Thus, ULK1-mediated autophagy is a mechanism by which the host resists the invasion of viruses or pathogens. Nevertheless, some viruses or pathogens can also employ the autophagy-related proteins or products of host cells (including ULK1) to promote their replication. In addition, ULK1 can directly mediate some proteins to resist infection caused by viral invasion by non-autophagic pathways.

4.4. Cardiovascular diseases

4.4.1. Autophagic function

Autophagy is an important way for most cardiovascular-derived cells to maintain homeostasis and is a necessary condition for regulating cardiovascular function258,259. Most heart-related diseases, including cardiovascular diseases, myocardial ischemia (IR), are generally accompanied by abnormal autophagy260, which might be caused by mitochondrial dysfunction and oxidative stress160.

It is worth noting that autophagy has complex functions in IR. Lipocalin-2, a protein positively associated with heart failure, was found to downregulate the phosphorylation of AMPK and ULK1 to reduce autophagic flux, inhibit beneficial autophagy during IR, and increase ischemia-induced cardiomyocytes death261. Notably, excessive autophagy aggravates IR damage. Currently, inhibiting autophagy by activating the Akt/mTOR/ULK1 pathway and inhibiting oxidative stress and cardiomyocyte apoptosis contributes to cardiomyocyte survival262. Intriguingly, there is a dramatic down-regulation of autophagy and levels of ULK1 in obesity-related cardiac dysfunction. High-fat diet feeding has no effect on triglyceride or diacylglycerol levels or cardiac function when Ulk1 and lipoprotein lipase are knocked out in mice. Therefore, obesity-induced cardiomyopathy can be treated by activating ULK1-mediated autophagy to clear proteolysis263. For heart damage caused by certain drugs, ULK1-mediated autophagy could always protect the heart. For example, the beneficial effect of isiduniol on DOX-induced myocardial damage depends on AMPK and ULK1 phosphorylation264. AMPK/mTOR/ULK1-mediated autophagy activation could slow disease progression in a model of isoproterenol-induced myocardial fibrosis265. In glucose-induced myocardial injury, the autophagy induction of mTOR/ULK1 plays a pivotal role. Additionally, experiments show that high glucose can increase ROS, phosphorylation of mTOR and its phosphorylation of ULK1 in H9c2 cardiomyocytes, further inhibiting autophagy and inducing apoptosis. Thus, mTOR-induced inactivation of ULK1 inhibits autophagy and protects cardiomyocytes from high glucose toxicity266. However, phosphorylation of ULK1 is not modulated by acute hyperglycemia267. Acute ethanol toxicity could activate AMPK and promote autophagy, triggering myocardial contractile dysfunction. Of note, this process is achieved by inhibiting mTORC1 and ULK1 phosphorylation268. In the abnormality of myocardial contraction induced by low ambient temperature, the phosphorylation of the autophagy-inhibiting signaling molecules Akt and mTOR decreases, while the phosphorylation of ULK1 increases, thereby causing an increase in cardiac autophagy269. Interestingly, long-term moderate exercise can produce benign changes in myocardial autophagy and protect the myocardium. Some scholars270 conducted long-term endurance exercise intervention on 8-week-old BALB/c mice. The experimental results showed that exercise training could significantly increase the expression and phosphorylation level of LC3II, Atg12, p62, and ULK1 (Ser555) proteins, protect cardiomyocytes, and maintain cardiac function. The AMPK/mTOR/ULK1 pathway is more critical in exercise-regulated autophagy. Different forms of exercise have various regulations on autophagy pathways. For example, endurance exercise can activate AMPK signaling pathway, inhibit mTOR, activate ULK1, and finally trigger autophagy271. In response to cardiac pressure overload, ULK1 also participates the initiation of mitophagy, thereby protecting the heart from cardiac insufficiency272. Furthermore, ULK1-Rab9-mediated mitophagy was activated during the chronic phase of high-fat diet depletion and rescued obese cardiomyopathy by controlling mitochondrial mass273.

4.4.2. Non-autophagic function

More recently, the non-autophagic role of ULK1 may affect heart-related diseases, which enriches its broader role beyond autophagy in cellular signaling pathways. Recent studies have found that deletion of ULK1 in the adult heart accelerates the progression of cardiomyopathy and heart failure, interestingly not by affecting the ULK1-mediated autophagy pathway274. In addition, ROS can affect the progression of various cardiac diseases, including cardiac hypertrophy, myocardial contractility disorders, cardiac extracellular matrix remodeling, and arrhythmias. Therefore, the reduction of oxidative stress is probably an effective preventive measure against stress overload-induced cardiac hypertrophy. Intriguingly, studies have shown that ULK1 levels are elevated in dilated cardiomyopathy. ULK1 can regulate ROS production by promoting the Nrf-2/HO-1 pathway, affecting angiotensin II-induced cardiac hypertrophy progression275.

In short, ULK1-mediated autophagy activation is beneficial to the heart in most cases, but excessive autophagy can aggravate the damage of the heart. In addition, ULK1 can also induce the production of ROS and induce cardiac hypertrophy. Therefore, due to the complex role of ULK1 in heart disease, there is still a long way to go to apply ULK1-related therapies to improve heart disease.

4.5. Other human diseases

4.5.1. Autophagic function

ULK1 also plays a unique role in some other organs, including the liver, bone, skin, etc. Associated diseases in these organs are generally accompanied by uncontrolled autophagy. For instance, SQSTM1/p62 activates NFE2L2/NRF2 and protects mouse liver against lipotoxicity through ULK1-mediated degradation of KEAP1 in autophagy145. Moderate exercise and healthy diet can promote lipophagy and improve non-alcoholic fatty liver disease and liver aging caused by a high-fat diet by activating AMPK/ULK1 and inhibiting Akt/mTOR/ULK1 pathway, respectively276. Intriguingly, acupuncture, as traditional Chinese medicine therapy and alternative therapy, can prevent and treat diseases by regulating autophagy level Electroacupuncture “Zusanli” can improve gastrointestinal motility disorder in rats with functional dyspepsia. Its mechanism may inhibit the excessive autophagy level of Cajal stromal cells by regulating the AMPK/ULK1 signal pathway. Unfortunately, the current research lacks unified conclusions and needs in-depth research277. In addition, ULK1/LC3 is inversely associated with renal fibrosis in diabetic patients, and dysregulation of autophagy leads to diabetic kidney disease, and activation of autophagy via p53/miR-214/ULK1 axis may alleviate diabetic kidney disease72. Besides, by activating the p38MAPK/mTOR/ULK1Ser757 signaling pathway, sPLA2-IB downregulates autophagy through PLA2R and causes podocyte damage, which in turn impairs glomerular filtration278. Furthermore, ULK1 expression is downregulated during osteoclast differentiation, and ULK1 may regulate osteoclasts differentiation through DOK3/Syk/JNK signaling in vitro, which is clinically relevant to osteoporosis. Therefore, activation of ULK1 is likely to be a promising therapeutics for osteoporosis279. Indeed, osteoprotegerin could inhibit osteoclast bone resorption by Akt/mTOR/ULK1-induced autophagy280. Notably, ULK1 has also been linked to skin diseases. In psoriasis, AMPK regulates autophagy through the ULK1/Atg7 signaling pathway and mitophagy through the PINK1/Parkin signaling pathway, thereby affecting the prognosis of a mouse model of psoriasis. Also, UVB irradiation can down-regulate the expression of ULK1 in human keratinocytes and reduce autophagy281. SREBP-1c blocks ULK1 sulfuration-dependent activation by reducing the activity of cystathionine γ-lyase, resulting in decreased autophagic flux, which may be responsible for hepatic steatosis130.

4.5.2. Non-autophagic function

Importantly, non-autophagic function of ULK1 also contributes a lot in human disease. AMPK-activated ULK1 can help embryonic stem cells maintain self-renewal and maintain versatility106. ULK1 can phosphorylate multiple sites of DSH in Wnt signaling pathway, which may affect Wnt signaling173. Besides, ULK1 phosphorylates Ser843 in the binding region of mineralocorticoid receptor ligand in renal intercalated cells, to maintain water salt balance282. In addition, the activation of mTORC1–ULK1–Atg13 pathway induced by rapamycin may alleviate the inflammatory response of rat model of chronic non-bacterial prostatitis283. Furthermore, liver-specific ULK1 knockout can reduce the liver injury caused by acetaminophen284. And p62 and ULK1 can be phosphorylated by PERK to protect the liver from lipotoxicity by activating the KEAP1–Nrf2 pathway146. Additionally, ULK1-regulated stearoyl-CoA desaturase 1 transcriptional regulation is associated with lipid metabolism in hepatocytes. Mechanistically, ULK1 increases transcription of stearoyl-CoA desaturase 1, reduces lipid droplet formation in the SFA model and alleviates lipotoxicity in hepatocytes285. Further, ULK1 can also protects cells by controlling RIPK1-mediated cell death286.

In general, ULK1 plays a complex and diverse role in a variety of human diseases. Due to the Jannus role of autophagy, the autophagy initiator ULK1 is inherently complicated, and its non-autophagy-related function also makes its role impossible to simply characterize. Therefore, the role of ULK1 needs to be further explored for ULK1-related therapeutics to have a promise for future potential therapeutic purpose.

5. Targeting ULK1 with pharmacological small-molecule compounds

More recently, some small-molecule compounds have been reported to affect disease progression and achieve therapeutic effects by modulating ULK1 (Fig. 7 and Table 4, Table 5, Table 6, Table 7). In this section, we focus on discussing some representative pharmacological small-molecule compounds that modulate the ULK1-mediated autophagy in different types of human diseases, analyzing their characteristics, advantages, and disadvantages, and looking forward to providing a clue on the development of ULK1-related drugs.

Figure 7.

Figure 7

Small-molecule compounds modulating the ULK1 pathways. ULK1 activators and inhibitors are roughly divided into two categories, one of which is directly affecting ULK1, and the other one is indirectly affecting ULK1 activity via regulating ULK1 pathway-related proteins. The chemical structures of ULK1 inhibitors, activators, and compounds that indirectly modulate ULK1 are included.

Table 4.

Small-molecule compounds that directly regulate ULK1 protein.

Compd. Structure Activator/inhibitor Target Disease ULK1
Activity
Biological Activity Ref.
Compound 6 Image 1 Inhibitor ULK1 / IC50 = 8 nmol/L Cell type:/
Animal type:/
16
XST-14 Image 2 Inhibitor ULK1 Hepatocellular carcinoma ULK1 (IC50 = 26.6 nmol/L) Cell type: Hep3B, HepG2 HCC cells, and L02 hepatocyte cells
Animal type: nude mice. Effective dose 15 or 30 mg/kg
214
MRT68921 Image 3 Inhibitor ULK1/2 Cancer ULK1 = 2.9 nmol/L, ULK2 = 1.1 nmol/L Cell type: human cancer cell lines A549, H1299, NCI-H460, MNK45, U251, SW480, SW620, HCT116, Colo320 and HT-29, PC-3, U266
Animal type: mice. Effective dose: 20 mg/kg
216,288,289
SBI-0206965 Image 4 Inhibitor ULK1/2 and AMPK cc RCC ULK1 (IC50 = 108 nmol/L); ULK2 (IC50 = 711 nmol/L) Cell type: A498 cells, ACHN cells
Animal type: C57BL/6NTac mice. Effective dose 10 μmol/L
222,292
Compound 3 Image 5 Inhibitor ULK1/2 / ULK1 (IC50 = 120 nmol/L) ULK2 (IC50 = 360 nmol/L) Cell type:/
Animal type:/
287
MRT67307 Image 6 Inhibitor ULK1/2 Cancer ULK1 = 45 nmol/L, ULK2 = 38 nmol/L Cell type: MEF cells
Animal type:/
288
5f Image 7 Inhibitor AURKA, FLT3, GSK3A, MAP3K, MEK, RSK2, RSK4, PLK4, ULK1, and JAK1 Lung cancer / Cell type: A549 NSCLC cells
Animal type:/
290
3g Image 8 Inhibitor ULK1 / ULK1(IC50 = 45 nmol/L) Cell type:/
Animal type:/
291
SBP-7455 Image 9 Inhibitor ULK1/2 TNBC ULK1(IC50 = 13 nmol/L) Cell type: MDA-MB-468, MDA-MB-231, and BT549 TNBC cells
Animal type: mice. Effective dose 10 mg/kg
293
3s Image 10 Inhibitor ULK1 NSCLC Anti-proliferative activity (IC50 = 1.94 ± 2.35 μmol/L) Cell type: A549 NSCLC cells Anti-proliferative activity (IC50 = 1.94 ± 2.35 μmol/L)
Animal type:/
294
LYN-1604 Image 11 Activator ULK1 TNBC EC50 = 18.94 nmol/L Cell type: MDA-MB-231 breast cancer cells. IC50 = 1.66 μmol/L
Animal type:/
15
BL-918 Image 12 Activator ULK1 PD EC50 = 24.14 nmol/L Cell type: SH-SY5Y cells, PC-12 cells.
Animal type: MPTP induced PD mouse model. Effective dose: 40 or 80 mg/kg
232

Table 5.

Small-molecule compounds that regulate the ULK1 pathways.

Compd. Structure Activation/
inhibition
Target Disease Biological activity Ref.
Nitazoxanide Image 13 Activation PI3K/AKT/mTOR/ULK1 and NQO1/mTOR/ULK1 pathways AD Cell type: BV2, SH-sy5y. Effective dose: 20 μmol/L
Animal type: APP/PS1 transgenic mice. Effective dose: 90 mg/kg
228
Chikusetsu saponin Iva Image 14 Activation AMPK/mTOR/ULK1 pathway Cardiac fibrosis Cell type:/
Animal type: BALB/c mice. 5 or 15 mg/kg
265
Ginkgolide K Image 15 Activation AMPK/mTOR/ULK1 pathway Ischemic stroke Cell type: astrocyte
Animal type:/
296
Melatonin Image 16 Activation AMPK/mTOR/ULK1 pathway Vascular calcification Cell type: vascular smooth muscle cells
Animal type:/
297
Salidroside Image 17 Activation AMPK/mTOR/ULK1 pathway Pulmonary hypertension Cell type: pulmonary arterial smooth muscle cells (PASMCs).
Animal type:/
298
Alisol A 24-acetate Image 18 Activation AMPK/mTOR/ULK1 pathway Nonalcoholic steatohepatitis Cell type: LX-2 human liver astrocyte cells.
Animal type: C57BL/6 mice. Effective dose: 15, 30 and 60 mg/kg
299
Berberine Image 19 Activation AMPK/mTOR/ULK1 pathway / Cell type: MPC5 podocytes.
Animal type:/
300
10-Hydroxycamptothecin (HCPT) Image 20 Activation AMPK/mTOR/ULK1 pathway Bladder cancer Cell type: T24 and 5637 human bladder cancer cells
Animal type:/
301
FL-411 Image 21 Activation AMPK/mTOR/ULK1 pathway TNBC Cell type: MDA-MB-231, MCF-7 human breast cancer cells. MCF-7 (IC50 = 1.62 μmol/L), MDA-MB-231 (IC50 = 3.27 μmol/L)
Animal type: female nude mice. Effective dose: 25, 50 and 100 mg/kg zebrafish. Effective dose: 12.5, 25 and 50 μmol/L.
302
Gossypol acetate Image 22 Activation AMPK/mTOR/ULK1 pathway Colon cancer, lung cancer Cell type: HCT116 cells and A549 cells. The best dose: 10 μmol/L 303
β-Guanidinopropionic acid Image 23 Activation AMPK/mTOR/ULK1 pathway / Cell type: Drosophila melanogaster. the effective dose to extend the lifespan of Drosophila melanogaster: ≥900 mmol/L
Animal type:/
304
Clozapine Image 24 Activation AMPK/ULK1 pathway Psychotic disorders Cell type:/
Animal type: Male Sprague–Dawley rats. Effective dose: 10 mg/kg
305
Eicosapentaenoic acid Image 25 Activation AMPK/ULK1 pathway Evaporative dry eye disease Cell type: human meibomian gland epithelial cells
Animal type:/
306
Ezetimibe Image 26 Activation AMPK/ULK1 pathway Middle cerebral artery occlusion Cell type:/
Animal type: Male Sprague–Dawley rats. Effective dose: 250 or 500 μg/kg, i.n. 5 or 10 mg/kg, i.p.
307
Rg2 Image 27 Activation AMPK/ULK1 pathway AD Cell type: HeLa human cervical carcinoma cell, Neuro2A Mouse brain neuroma cell, PC12 pheochromocytoma cell. The dose used in Hela cell: 50 nmol/L
Animal type: mice. Effective dose: 10 mg/kg and 20 mg/kg.
308
Resveratrol Image 28 Activation AMPK/ULK1 pathway / Cell type: mouse embryonic stem cells. the dose used in mouse embryonic stem cells: 10 μmol/L
Animal type:/
309
A77 1726 Image 29 Activation AMPK/ULK1 pathway ALS Cell type: NSC34 mouse neuron cells. The effective dose used in NSC34 cells: 200 μmol/L
Animal type:/
310
Kinsenoside Image 30 Activation AMPK/ULK1 pathway Alcoholic fatty liver Cell type: AML12 cells. The best dose: 40 μmol/L
Animal type: mice. 20 mg/kg
311
Baicalein Image 31 Activation AMPK/ULK1 pathway Human prostate and breast cancer Cell type: PC-3 human prostate cells, DU145 human prostate cells and MDA-MB-231 breast cancer cells. The effective dose used in DU145 cells: 2.5 μg/mL
Animal type:/
312
Narciclasine Image 32 Activation AMPK/ULK1 pathway TNBC Cell type: HCC-1937 and MDA-MB-231 breast cancer cells. the dose used in the HCC1937 cells: 20 or 50 nmol/L
Animal type: BALB/c nude mice. Effective dose: 2 or 5 mg/kg
314
Raloxifene Image 33 Activation AMPK/ULK1 pathway Breast cancer Cell type: MCF-7 human breast cancer cells. MCF-7(IC50 48 h = 10 μmol/L)
Animal type:/
315
3,3′-Diindolylmethane Image 34 Activation AMPK/ULK1 pathway Human prostate cancer Cell type: LNCaP and C42B human prostate cancer cells. the dose used in C42B cells: 30 μmol/L; the dose used in LNCaP cells: 20 μmol/L
Animal type:/
316
Temozolomide Image 35 Activation ATM/AMPK/ULK1 pathway Glioma Cell type: U87MG and U251 human malignant glioma cells. the dose induced autophagy: 100 μmol/L
Animal type:/
317
Aescin Image 36 Activation ATM/AMPK/ULK1 pathway HCC and colon carcinoma Cell type: HepG2 HCC cells and HCT 116 colon carcinoma cells. the dose used in HepG2 cells and HCT116 cells: 20–80 μg/mL
Animal type:/
318
Aspirin Image 37 Activation mTOR/ULK1 pathway Murine hepatocarcinoma and sarcoma Cell type: H22 hepatocarcinoma and S180 sarcoma cells
Animal type: male Kunming mice. Effective dose: 100 or 400 mg/kg
319
AZD8055 Image 38 Activation mTOR/ULK1 pathway / Cell type: HT-29, DLD-1 colon carcinoma cells.
Animal type:/
320
Polyphyllin VI Image 39 Activation AMPK/mTOR/ULK1 pathway Non-small cell lung cancer Cell type: A549 NSCLC cells, and H1299 NSCLC cells.
Animal type: BALB/c nude mice. Effective dose: 2.5, 5 or 10 mg/kg
321
Isorhamnetin (IH) Image 40 Activation PI3K/AKT/mTOR/p70S6K/ULK1 pathway TNBC Cell type: MDA-MB-231 breast cancer cells. Inhibition of cell proliferation in MDA-MB-231: 55.51 μmol/L
Animal type:/
322
Genkwanin (GN) Image 41 Activation PI3K/AKT/mTOR/p70S6K/ULK1 pathway TNBC Cell type: MDA-MB-231 breast cancer cells. Inhibition of cell proliferation in MDA-MB-231: 58.54 μmol/L
Animal type:/
322
Acacetin Image 42 Inhibition PI3K/AKT/mTOR/p70S6K/ULK1 pathway TNBC Cell type: MDA-MB-231 breast cancer cells. Inhibition of cell proliferation in MDA-MB-231: 82.75 μmol/L
Animal type:/
322
Niclosamide Image 43 Inhibition mTOR/ULK1 pathway Ischemic stroke Cell type: HEK293T human embryonic kidney cells. HCT116 colon cancer cell (IC50 = 0.31 μmol/L)
Animal type:/
323
Tizoxanide Image 44 Activation PI3K/Akt/mTOR/ULK1 pathway Virus Cell type: HepG2 HCC cells.
Animal type:/
324
LicA Image 45 Activation ULK1/Atg13 and ROS pathway HCC Cell type: HuH7 and HepG2 HCCs cells. The best dose: 50 μmol/L
Animal type:/
325
4,4′-Dimethoxychalcone Image 46 Activation p38 and JNK pathway Aging Cell type: HaCaT cells. The best dose: 20 μmol/L
Animal type: mice. Effective dose: 50, 100 mg/kg
326
20-Hydroxyecdysone Image 47 Activation Atg genes / Cell type: silkworm. The dose used in HEK 293 cells: 1 μmol/L
Animal type: the silkworms. Effective dose: 5 μg/larva
327
Vitexin Image 48 Inhibition mTOR/ULK1 pathway Cerebral ischemic stroke Cell type:/
Animal type: MACO model SD rats. Effective dose: 2 mg/kg
328
WP1130 Image 49 Inhibition ULK1, USP9X Cervical cancer and osteosarcoma Cell type: HeLa cervical cancer cells U20S human osteosarcoma cells
Animal type:/
329
Geldanamycin Image 50 Inhibition Atg7, Beclin-1, ULK1 The cytotoxicity of sunitinib in cardiomyocytes Cell type: NH9c2 rat myocardial cells.
Animal type:/
331
Hyperoside Image 51 Inhibition AMPK/ULK1 pathway Renal aging Cell type: NRK-52E rat kidney cells
Animal type:/
332
Isoliquiritigenin Image 52 Inhibition AMPK/mTOR/ULK1 pathway TNBC Cell type: MDA-MB-231 breast cancer cell
Animal type: female Nude-Foxn1 mice. Effective dose: 2.5 or 5 mg/kg
333

Table 6.

Drug combination therapy.

Compd. 1 Structure 1 Compd. 2 Structure 2 Activation/Inhibition Target Cell type Disease Ref.
Curcumin Image 53 5-FU Image 60 Inhibition AMPK/ULK1 autophagy HCT116 and HT29 colon cancer cells Colon cancer 208
Chloroquine Image 54 5-FU Image 61 Activation ULK1 HCT-116 colon cancer cells Colon cancer 334
Enzalutamide Image 55 3-MA Image 62 Inhibition AMPK/mTOR/ULK1 pathway J82, T24, and UMUC3 human bladder cancer cells Bladder cancer 335
Enzalutamide Image 56 BAF Image 63 Inhibition AMPK/mTOR/ULK1 pathway J82, T24, and UMUC3 human bladder cancer cells Bladder cancer 335
Enzalutamide Image 57 CQ Image 64 Inhibition AMPK/mTOR/ULK1 pathway J82, T24, and UMUC3 human bladder cancer cells Bladder cancer 335
AZD5363 Image 58 FH535 Image 65 Activation AMPK/mTOR/ULK1 pathway HepG2 HCC cells and Hep3B cells Hepatocellular carcinoma 336
Ascorbic acid Image 59 Menadione Image 66 Activation AMPK/mTORC1/ULK1 pathway Glioblastoma Glioblastoma Cells 337

Table 7.

Other therapeutic strategies.

Drug Activation/
Inhibition
Target Cell type Disease Ref.
Heavy metal scavenger metallothionein mitigates Activation Phosphorylation of ULK1 / Myocardial contractile anomalies 269
Extracellular histones Activation Sestrin2/AMPK/ULK1–mTOR and AKT/mTOR Human endothelial cell / 338
The bursa of Fabricius (BP7) Activation AMPK–ULK1 phosphorylation WEHI-231 mouse B lymphocytes cells / 339
Adiponectin Activation LC3, Beclin1, ULK1 L6 skeletal muscle cells Diabetes 339

5.1. Small-molecule inhibitors

In 2015, based upon the crystallographic structure of ULK1 kinase domain, combined with kinase activity assay, the first ULK1 inhibitor compound 6 was designed based on the structure of ULK1 by Lazarus, with ULK1 IC50 = 8 nmol/L. Unfortunately, it is not a specific inhibitor for ULK116. Subsequently, a new scaffolding ULK1 inhibitor compound 3 was found to be a useful tool compound, whose discovery further deepened the understanding of the structure of ULK1 inhibitors, and it was higher selectivity to ULK1, while lower affinity compared to compound 6 (ULK1 IC50 = 120 ± 1.7 nmol/L, ULK2 IC50 = 360 ± 79 nmol/L)287. Another two ULK1 inhibitors, MRT67307 and MRT68921, showed excellent inhibitory effects on both ULK1 and ULK2, among which the IC50 of MRT67307 for ULK1 is 45 nmol/L, and the IC50 for ULK2 is 38 nmol/L, and MRT68921 exhibited better activity (IC50 for ULK1 = 2.9 nmol/L, IC50 for ULK2 = 1.1 nmol/L)288. Of note, MRT68921 displayed good anti-tumor activity in vitro and in vivo289. In addition, the tubulin polymerization inhibitor 5f was found to inhibit the enzymatic activity of various kinases, including ULK1. Although 5f has tumor-suppressive potential, its high cytotoxicity limits its use290. Besides, the ULK1 small molecule inhibitor 3g with an indazole core was discovered through high-throughput in silico screening (HTS) followed by structure optimization. This compound was found to have a high selection to inhibit ULK1 (IC50 = 45 nmol/L) with outstanding stability in vivo and good inhibition of CYP291. Interestingly, Egan and his colleagues also discovered another famous ULK1 kinase inhibitor SBI-0206965, with higher selection, with IC50 values of 108 nmol/L for ULK1 and 711 nmol/L for ULK2, which is the most used ULK1 inhibitor. In addition, SBI-0206965 could suppress ULK1-mediated phosphorylation events, inhibiting autophagy14. In addition, SBI-0206965 could down-regulate ULK1 and selectively inhibit ULK1-induced apoptosis in renal cell carcinoma cells223,292. Compared with SBI-0206965, SBP-7455 has increased binding affinity for ULK1/2 (IC50 for ULK1 = 13 ± 2 nmol/L, IC50 for ULK2 = 476 ± 21 nmol/L), potently inhibits ULK1/2 activity in vitro, and reduces TNBC cell proliferation, and is orally bioavailable in vivo. SBP-7455 inhibits autophagy for autophagy-dependent survival in TNBC cells even when starved and exhibits synergistic cytotoxicity against TNBC cells with PARP inhibitor olaparib293. According to the characteristics of the interaction between ULK1 kinase and ULK1 inhibitors, the study of the features of chemical structures, as well as high-throughput virtual screening, 3s was found to be a good ULK1 inhibitor with 99.15% ULK1 kinase inhibitory activity at 10 μmol/L, higher than 97.56% of SBI-0206965. And 3s also exert a promising effect of anti-non-small cell lung cancer294. Additionally, ULK1 inhibitor XST-14 was discovered through structure-based virtual docking (IC50 value for ULK1 is 13.6 nmol/L), which acts as an anti-hepatoma drug by reducing autophagy and inducing apoptosis214. The above-mentioned ULK1-targeted inhibitors and their related mechanisms of action and applications are summarized in Table 4. It is worth noting that although a variety of ULK1 inhibitors have been discovered one after another with the modernization of medicinal chemistry, unfortunately, none of them can be called drugs in the true sense, nor have they entered clinical trials. Nevertheless, ULK1 inhibitors still have potential applications in human disease. Particularly, MRT68921, SBI-0206965, SBP-7455, and XST-14 all showed good anti-tumor effects in animal experiments, which also provided the possibility for their further development.

5.2. Small-molecule activators

Through TCGA analysis, TMA analysis, as well as in silico screening and chemical synthesis, LYN-1604 was discovered as a good candidate as a ULK1 activator and an excellent agent to treat TNBC by regulating autophagy. Notably, LYN-1604 is the first ULK1 activator with half effective concentration for activating ULK1 is 18.94 nmol/L, and showed good anti-TNBC activity in vitro and in vivo. Therefore, LYN-1604 may have therapeutic potential in future TNBC treatment15. Recently, another ULK1 activator BL-918 was found through structure-based drug design. BL-918 has an EC50 = 24.14 nmol/L for ULK1, and could rescue SH-SY5Y cells from MPP+ injury, and protects motor disorder and dopaminergic neuron loss induced by MPTP in mouse by activating ULK1-regulated autophagy (Table 4)232. Invigoratingly, the ULK1 activator to regulate autophagy holds promise for physiological and pharmacological perspectives for treating human diseases.

5.3. Small-molecule compounds modulating the ULK1-related pathways

5.3.1. Positive regulation of ULK1

As the most commonly canonical autophagic pathway, the AMPK/mTOR/ULK1 pathway affects multiple diseases by activating autophagy295. Correspondingly, several compounds could positive regulative ULK1 via this pathway. For instance, ginkgolide K296 promotes protective autophagy to promote astrocyte proliferation and migration after oxygen-glucose deprivation to treat ischemic stroke; Chikusetsu saponin Iva275 could relieve disease progression by increasing autophagic activity in isoproterenol-modeled mouse myocardial fibrosis. Melatonin297 mitigates vascular calcification through activation of autophagy. Importantly, AMPK/mTOR/ULK1 pathway also affects energy metabolism. In hypoxia, salidroside reduces pulmonary artery smooth muscle cells activity and its resistance to apoptosis by upregulating autophagy298. Alisol A 24-acetate inhibits oxidative stress and activates autophagy in mouse liver and LX-2 cells, which contribute to the treatment of nonalcoholic steatohepatitis299. Under high glucose conditions, berberine protects podocytes from damage by activating autophagy300. In addition, 10-hydroxycamptothecin301, FL-411302, gossypol acetate303 serve as anticancer agents by improving autophagic activity mediated by AMPK/mTOR/ULK1 pathway. Interestingly, in Drosophila, β-guanidinopropionic acid can contribute to prolonging lifespan by AMPK/Atg1 pathway304. Clozapine can stimulate the AMPK–ULK1–Beclin1 pathway by inducing the phosphorylation of ULK1 and Beclin1 and enhancing the levels of LC3-II and Atg5–Atg12, thus promoting autophagic flux to eliminate misfolded proteins in neural cells and degrade toxic substances305. Eicosapentaenoic acid could exert neuroprotective effects after MCAO by inducing AMPK/ULK1 signal and autophagy activation through PPARγ signal306. Ezetimibe is an NPC1L1 inhibitor that activates the AMPK/ULK1/autophagy pathway by phosphorylating AMPK, ULK1, Beclin1, etc., reducing apoptosis of neurocyte, and is worthy of intensive study as a promising agent for treating ischemic stroke307. Additionally, compound Rg2 treats metabolic diseases by activating the AMPK/ULK1 autophagy pathway, such as the clearance of protein aggregates, high-fat diet-induced insulin resistance, etc308. Activation of the AMPK/ULK1 pathway by resveratrol could control pluripotency of embryonic stem cells in mice309. Compound A77 1726 inhibits S6K1 and promotes autophagy by increasing phosphorylation of AMPKThr172 and ULK1Ser555, and enhances degradation of SOD1G93A protein for the treatment of ALS310. Kinsenoside activates AMPK–ULK1-dependent autophagy to alleviate alcoholic liver injury311. Activation of autophagy also plays a therapeutic role in cancer therapy regulated by AMPK/ULK1 signaling axis, such as baicalein against multiple cancers312,313; narciclasine against TNBC314; raloxifene315 against breast cancer; 3,3′-diindolylmethane against human PC316. As for drug-resistant, temozolomide activates autophagic activity through ATM–AMPK–ULK1 pathway, leading to chemotherapy resistance in glioma317. Also, aescin induces ROS activation and promotes autophagy through activation of the ATM/AMPK/ULK1 pathway, which might promote cancer cell survival; thus, combined use with autophagy inhibitors may provide a therapeutic strategy for anti-tumor318. Aspirin may induce autophagy by inhibiting mTOR signaling targets and increasing ULK1 and LC3A, thereby inhibiting angiogenesis in mouse liver cancer and sarcoma models319.

Additionally, the mTOR inhibitor AZD8055 inhibits phosphorylation of ULK1 at Ser757 to activate ULK1 activity and induces autophagic activity and downregulates p62 to antagonize chemotherapy-induced cell death320. Polyphyllin VI promotes apoptosis and autophagic cell death in NSCLC through ROS-induced mTOR/ULK1 pathway and exerts potent anti-cancer proliferation in vitro and in vivo in its prototype form321. In human breast cancer cells, flavonoids such as isorhamnetin, coriander, and acacetin could downregulate PI3Kγ-p110 and subsequently disrupt the PI3K/Akt/mTOR/p70S6K/ULK1 pathway, then inhibiting cancer cell proliferation by inducing cell cycle arrest at the G2/M phase of breast cancer cells, promoting apoptosis and autophagic cell death322. In niclosamide-resistant cell lines, mTOR/ULK1 signaling, was blocked, and LC3B expression was increased, suggesting that autophagy may be a marker of niclosamide tumor resistance323. Tizoxanide may enhance autophagic activity by blocking the Akt/mTOR/ULK1 signaling pathway in macrophages324. In AD model mice, nitazoxanide, an anti-parasitic drug, could activate autophagic flux through PI3K/Akt/mTOR/ULK1 and NQO1/mTOR/ULK1 signaling pathways blockage to clear Aβ protein in the brain, which is beneficial to ameliorate learning and memory impairments228.

Further, licochalcone A (LicA), a novel chemotherapeutic agent, as a Bcl-2 inhibitor, which can induce autophagy in human HCC cells via ULK1/Atg13 and ROS pathways, and inhibition of LicA-induced ROS by the antioxidant NAC can enhance LicA-induced cells apoptosis, further inhibiting HCC325. 4,4′-Dimethoxychalcone can affect ULK1 to activate autophagy through JNK and MAPK kinases slowing skin aging326. And 20-hydroxyecdysone can up-regulate Atg gene and induce autophagy in the silkworm fat body (Table 5)327.

5.3.2. Negative regulation of ULK1

Some small molecular compounds may inhibit autophagy by inhibiting ULK1-related signaling pathways to treat diseases (Table 5). Vitexin inhibits autophagy dysfunction through mTOR/ULK1 pathway to alleviate MCAO-induced oxidative damage, apoptosis and inflammation, thereby improving cerebral ischemic stroke328. Inhibition of ULK1 by WP1130 blocks autophagic flux through inhibition of deubiquitinase, enhancing ULK1 ubiquitination and ULK1 transfer to aggregates329. Rocaglamide inhibits autophagy by inhibiting ULK1 at the protein translation level, a process that restores granzyme B levels in NK cells to increase NK cell sensitivity to NSCLC in cancer immunotherapy330. Sunitinib-induced myocardial cytotoxicity is a severe side effect during treatment; and studies have found that the HSP90 inhibitor geldanamycin can reduce the side effect by degrading autophagy-related proteins such as ULK1 and inhibiting the autophagy pathway331. Hyperoside inhibits AMPK/ULK1-mediated autophagy in a rat kidney injury model following D-gal treatment, thereby preventing age-related kidney injury332. In addition, Dietary compound isoliquiritigenin reduces the total expression and phosphorylation of ULK1 in mammals and inhibits the expansion of TNBC cells through autophagy-mediated apoptosis333.

5.4. Drug combination therapies

On account of the complex role of autophagy in diseases, drug combination therapy is increasingly recognized for better long-term prognosis and fewer side effects, especially when drug resistance occurs. The combination of autophagy inhibitor chloroquine and low concentration 5-FU can increase the ability of tumor cells to induce dendritic cell (DC) maturation by activating ULK1334. Importantly, ULK1-mediated autophagy may be one of the reasons for the resistance in enzalutamide therapy in bladder cancer. Therefore, enzalutamide concurrent treatment with autophagy inhibitors (including 3-MA, BAF, and chloroquine) might be an emerging strategy to treat bladder cancer effectively335. The combination of β-catenin inhibitor (FH535) and Akt inhibitor (AZD5363) enhances p53 expression and further induces lethal autophagy via regulation of the AMPK–mTOR–ULK1 pathway, exerting a more substantial effect on cell death of transformed hepatocytes, playing a treatment role in human hepatoma336. Combination of ascorbic acid and menadione also exerts anticancer effects through the induction of cytotoxic autophagy in human glioblastoma through the AMPK/mTOR/ULK1 pathway337. Curcumin pretreatment combined with 5-Fu enhanced the sensitivity of 5-Fu to colon cancer cells by inhibiting AMPK/ULK1-dependent autophagy (Table 6)208.

5.5. Other therapeutic strategies

Many protein drugs can also affect the expression of ULK1. Extracellular histones could promote autophagic and apoptotic activity through Sestrin2/AMPK/ULK1/mTOR and Akt/mTOR pathway to inhibit cultured human endothelial cells, which may be a novel target for therapeutic strategies in endothelial damage338. Heavy metal scavenger can increase the phosphorylation of ULK1 and inhibit the phosphorylation of eNOS, reducing the abnormal myocardial contraction caused by deep hypothermia by attenuation of cardiac autophagy269. A new bursal heptapeptide (BP7) stimulates autophagy and phosphorylation of AMPK/ULK1 pathway and regulates the expression of Bcl-2 protein in WEHI-231 cells hence acting as an active biological factor and can be used as a potential immune enhancer339. In mice, adiponectin stimulation, could correct HFD-induced LC3, Beclin1, and Ulk1 gene expression, reduces autophagy, and reduces oxidative stress, thereby increasing insulin sensitivity (Table 7)340.

6. Conclusions and perspectives

ULK1 is a key regulator in the initiation of mammalian autophagy, and its function is conserved in all eukaryotes. Accordingly, ULK1 has a potential to fight human diseases as a promising therapeutic target. With the in-depth studies of ULK1, whether it is the canonical autophagy reaction played by ULK1 by autophagy–lysosomal pathway, or its independent of autophagy role in stabilizing oxidative stress and endoplasmic reticulum stress, reducing lipid, promoting erythrocyte maturation, as well as regulating innate immune response, shows that ULK1 plays an irreplaceable role in normal physiological activities and regulating pathological processes.

Moreover, autophagy may be beneficial or detrimental depending on the type or stage of one disease. For example, ULK1 plays the Janus role in different types and stages of human cancers. According to the specific development stages and types of a particular tumor, developing small-molecule drugs targeting ULK1 to impede the tumor progression is a worthy researching direction. For neurodegenerative diseases, ULK1 mainly plays a protective role by activating autophagy. But in a few neurodegenerative diseases such as Chorea-acanthocytosis, ULK1 can play a pathogenic role. The specific role of ULK1 in various nervous system diseases still needs to be explored by further deepening study. In immune diseases, whether ULK1-mediated autophagy acts to fight viral invasion or promote viral replication depends on the type of virus. ULK1-mediated moderate autophagy activation is beneficial to the heart, but excessive autophagy can aggravate the process of heart disease. The precise correction of the uncontrolled ULK1 is appropriate. In addition, there are few studies on the unique role of ULK1 in tissues and organs, such as skin, kidneys, bone, and liver. Thus, whether ULK1 plays a therapeutic role or a toxic reaction is also worthy of further investigations. Interestingly, ULK1 is explicitly expressed in some organs or lesions, facilitating the pathological process via non-autophagy pathways. For example, the anti-migration effect of ULK1 phosphorylation of Exo70 in tumors is not achieved through the autophagy pathway. Due to its structural characteristics, ULK1 plays a key role in regulating neuronal development. In addition, ULK1 promotion can directly affect the antiviral effect of IFN-γ. ULK1 in the adult hearts treats cardiomyopathy and heart failure through a non-autophagic pathway. As mentioned above, the non-autophagic role of ULK1 is also crucial in human diseases, beyond its inherent autophagy.

Since ULK1 is destined to have multifaceted roles in diverse human diseases, developing therapeutic approaches that integrate multiple pathways may be a successful therapeutic strategy. In some cases, the improvement of ULK1 for disease is based on its autophagy regulation function. Under this circumstances, excessive inhibition/activation of ULK1 may have unpredictable pathogenic effects on the other organ and tissue by affecting autophagy. Thus, specific regulation of target organ ULK1 may address this issue, and correction of abnormal modifications of disease-specific ULK1 may also avoid the damage to other tissues. In addition, more studies need to elucidate the precise function of ULK1-mediated autophagy in individual disease types and pathological processes before ULK1-targeted therapy can be considered. In other cases, ULK1 may ameliorate the disease through its non-autophagy function. At this time, the regulation of ULK1 on autophagy may bring risks to the treatment of the disease. In this case, a combination of multiple treatment regimens may be more appropriate (e.g., combined with autophagy inhibitors to correct autophagy levels under certain circumstances). Additionally, considering that the inhibition/activation of ULK1 appears to be limited by differences in ULK1 expression levels and basal values of autophagy in different types of organs/lesions, therapeutic strategies targeting ULK1 may have significant limitations. Thus, the combination of mechanism-based ULK1 inhibitors/activators with chemotherapeutics or targeted drugs may be more practical. Notably, the development of specific novel ULK1-targeting small-molecule drugs (targeting specific sites, specific modifications, specific PPI networks, etc.) may provide some new possibilities for potential applications of targeting ULK1 in human disease.

Unfortunately, the current research on inhibitors and activators targeting ULK1 to treat diseases is in the preclinical stage since the discovery of ULK1 inhibitors/activators is still in its infancy. It's exciting that another autophagy-related target, mTOR, is more mature in clinical research. For example, the mTOR inhibitor sirolimus combined with metformin is already in Phase 1 trials in advanced solid tumors. (NCT02145559)341. Notably, autophagy contributes a lot in this trial. Compared with mTOR, ULK1 has obvious autophagy induction ability, and the clinical application of drugs targeting ULK1 is extremely promising. There's optimism that some small molecular compounds targeting ULK1 exhibited good activity in vivo. For example, the ULK1 inhibitor SBI-0206965 and the ULK1 activator LYN-1604 showed excellent anticancer effects in nude mice, and the ULK1 inhibitor XST-14 displayed good pharmacokinetics in mice. As mentioned above, the clinical performance of ULK1 inhibitors/activators is worth for further expectations.

Currently, multiple issues and challenges have remained in translating knowledge on ULK1 into the clinic. First, the applications of small-molecule compounds targeting ULK1 should be attributed to a comprehensive understanding of its mechanisms of action in vitro and in vivo, but only part of it has been well elucidated. For example, the crystal structures of ULK1 and ULK1 complex have not been fully resolved, and increasing research in this area will help the discovery of small-molecule compounds that specifically target ULK1 and be further used for potential treatment. Further, potential toxic and side effects are the critical issue hindering the clinical administration of drugs. Notably, in the era of precision medicine, whether the non-autophagic function of ULK1 could guide the stratification of patients in clinical trials by determining ULK1 as a biomarker, which will facilitate the development of gene drugs, RNAi drugs, and antibody drugs targeting ULK1. Additionally, the functions of ULK1 in non-autophagy can contribute to predicting the most likely outcome for the patients who would benefit from ULK1 inhibitors/activators in a clinical setting. Therefore, elucidating the role of the non-autophagic function of ULK1 in various diseases is helpful for the development of drugs targeting ULK1. In general, addressing the above questions requires an in-depth understanding of the molecular mechanisms of ULK1 in some regulatory events, which may provide the possibility for the clinical and applications of ULK1-targeted therapy. The excitement that there are some novel methods to discover further the relations of the biological functions of ULK1 and human diseases and to find novel small molecular compounds for activating or inhibiting ULK1, such as CRISPR/Cas9, PRATACs, and artificial intelligence/machine learning. Pioneer studies have shown that disruption of the ULK1 gene can be easily knocked out or introduced by CRISPR/Cas9-induced dual fluorescent reporter genes342. PROTACs could effectively inactivate ULK1 protein function via compound-mediated proteolysis343. Artificial intelligence/machine learning is used to estimate autophagy protein levels. Autophagic proteins including ATG1, ATG5, and LC3B can be quantified by software-based integrated optical density to characterize alters in basal autophagy levels of patients344.

In summary, these inspiring findings demonstrate the molecular structure and biological functions of ULK1, elucidating its regulatory role in both autophagy and non-autophagy, which may ultimately provide a new clue on our understanding of the autophagic initiator ULK1. Moreover, we focus on discussing a comprehensive exploration of the complicated relationship between ULK1 and human diseases, establishing a specific theoretical basis for ULK1 for potential therapeutic purposes. The discovery of small-molecule drugs targeting ULK1 will become one of the dominant therapeutic strategies for fighting human diseases. More importantly, some new emerging technologies have still been expected to help discover the first-in-class drugs targeting ULK1; and discovery of more candidate small-molecule drugs would contribute to driving ULK1-targed therapy to the future clinic.

Acknowledgments

This work was supported in part by National Natural Science Foundation of China (Grant Nos. 82172649 and 82173666), Shenzhen science and technology research and development funds (Grant No. JCYJ20210324094612035, China), as well as the Key R&D Program of Sichuan Province (Grant No. 2021YFS0046, China). We also thank some materials in the graphical abstract and figures that are produced by Servier Medical Art (https://smart.servier.com) and BioRender (https://biorender.com).

Author contributions

Bo Liu and Jin Zhang conceived the project and supervised the project. Ling Zou, Minru Liao, Yongqi Zhen, and Shiou Zhu summed up the literature and drafted the manuscript. Xiya Chen was involved in chemical structure drawing. Ling Zou and Jin Zhang proofread the structure and figures. Bo Liu, Yue Hao, and Jin Zhang revised the manuscript. All authors approved the final manuscript.

Conflicts of interest

The authors have no conflicts of interest to declare.

Footnotes

Peer review under responsibility of Chinese Pharmaceutical Association and Institute of Materia Medica, Chinese Academy of Medical Sciences.

Contributor Information

Jin Zhang, Email: zhangjin1989@szu.edu.cn.

Yue Hao, Email: yuehao@szu.edu.cn.

Bo Liu, Email: liubo2400@163.com.

References

  • 1.Mizushima N., Komatsu M. Autophagy: renovation of cells and tissues. Cell. 2011;147:728–741. doi: 10.1016/j.cell.2011.10.026. [DOI] [PubMed] [Google Scholar]
  • 2.Byrnes K., Blessinger S., Bailey N.T., Scaife R., Liu G., Khambu B. Therapeutic regulation of autophagy in hepatic metabolism. Acta Pharm Sin B. 2022;12:33–49. doi: 10.1016/j.apsb.2021.07.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Marx J. Autophagy: is it cancer's friend or foe? Science. 2006;312:1160–1161. doi: 10.1126/science.312.5777.1160. [DOI] [PubMed] [Google Scholar]
  • 4.Ahsan A., Liu M., Zheng Y., Yan W., Pan L., Li Y., et al. Natural compounds modulate the autophagy with potential implication of stroke. Acta Pharm Sin B. 2021;11:1708–1720. doi: 10.1016/j.apsb.2020.10.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Tsukada M., Ohsumi Y. Isolation and characterization of autophagy-defective mutants of Saccharomyces cerevisiae. FEBS Lett. 1993;333:169–174. doi: 10.1016/0014-5793(93)80398-e. [DOI] [PubMed] [Google Scholar]
  • 6.Rabinowitz J.D., White E. Autophagy and metabolism. Science. 2010;330:1344–1348. doi: 10.1126/science.1193497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Kuroyanagi H., Yan J., Seki N., Yamanouchi Y., Suzuki Y., Takano T., et al. Human ULK1, a novel serine/threonine kinase related to UNC-51 kinase of Caenorhabditis elegans: cDNA cloning, expression, and chromosomal assignment. Genomics. 1998;51:76–85. doi: 10.1006/geno.1998.5340. [DOI] [PubMed] [Google Scholar]
  • 8.Hara T., Takamura A., Kishi C., Iemura S., Natsume T., Guan J.L., et al. FIP200, a ULK-interacting protein, is required for autophagosome formation in mammalian cells. J Cell Biol. 2008;181:497–510. doi: 10.1083/jcb.200712064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Ganley I.G., Lam du H., Wang J., Ding X., Chen S., Jiang X. ULK1.ATG13.FIP200 complex mediates mTOR signaling and is essential for autophagy. J Biol Chem. 2009;284:12297–12305. doi: 10.1074/jbc.M900573200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Chen S., Wang C., Yeo S., Liang C.C., Okamoto T., Sun S., et al. Distinct roles of autophagy-dependent and -independent functions of FIP200 revealed by generation and analysis of a mutant knock-in mouse model. Genes Dev. 2016;30:856–869. doi: 10.1101/gad.276428.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Zheng X., Li W., Xu H., Liu J., Ren L., Yang Y., et al. Sinomenine ester derivative inhibits glioblastoma by inducing mitochondria-dependent apoptosis and autophagy by PI3K/AKT/mTOR and AMPK/mTOR pathway. Acta Pharm Sin B. 2021;11:3465–3480. doi: 10.1016/j.apsb.2021.05.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Egan D.F., Shackelford D.B., Mihaylova M.M., Gelino S., Kohnz R.A., Mair W., et al. Phosphorylation of ULK1 (hATG1) by AMP-activated protein kinase connects energy sensing to mitophagy. Science. 2011;331:456–461. doi: 10.1126/science.1196371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Wang B., Kundu M. Canonical and noncanonical functions of ULK/Atg1. Curr Opin Cell Biol. 2017;45:47–54. doi: 10.1016/j.ceb.2017.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Egan D.F., Chun M.G., Vamos M., Zou H., Rong J., Miller C.J., et al. Small molecule inhibition of the autophagy kinase ULK1 and identification of ULK1 substrates. Mol Cell. 2015;59:285–297. doi: 10.1016/j.molcel.2015.05.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Zhang L., Fu L., Zhang S., Zhang J., Zhao Y., Zheng Y., et al. Discovery of a small molecule targeting ULK1-modulated cell death of triple negative breast cancer in vitro and in vivo. Chem Sci. 2017;8:2687–2701. doi: 10.1039/c6sc05368h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Lazarus M.B., Novotny C.J., Shokat K.M. Structure of the human autophagy initiating kinase ULK1 in complex with potent inhibitors. ACS Chem Biol. 2015;10:257–261. doi: 10.1021/cb500835z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Wong P.M., Puente C., Ganley I.G., Jiang X. The ULK1 complex: sensing nutrient signals for autophagy activation. Autophagy. 2013;9:124–137. doi: 10.4161/auto.23323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Hurley J.H., Young L.N. Mechanisms of autophagy initiation. Annu Rev Biochem. 2017;86:225–244. doi: 10.1146/annurev-biochem-061516-044820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Noda N.N., Fujioka Y. Atg1 family kinases in autophagy initiation. Cell Mol Life Sci. 2015;72:3083–3096. doi: 10.1007/s00018-015-1917-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Ogura K., Okada T., Mitani S., Gengyo-Ando K., Baillie D.L., Kohara Y., et al. Protein phosphatase 2A cooperates with the autophagy-related kinase UNC-51 to regulate axon guidance in Caenorhabditis elegans. Development. 2010;137:1657–1667. doi: 10.1242/dev.050708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Yan J., Kuroyanagi H., Kuroiwa A., Matsuda Y., Tokumitsu H., Tomoda T., et al. Identification of mouse ULK1, a novel protein kinase structurally related to C. elegans UNC-51. Biochem Biophys Res Commun. 1998;246:222–227. doi: 10.1006/bbrc.1998.8546. [DOI] [PubMed] [Google Scholar]
  • 22.Papinski D., Kraft C. Atg1 kinase organizes autophagosome formation by phosphorylating Atg9. Autophagy. 2014;10:1338–1340. doi: 10.4161/auto.28971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Toda H., Mochizuki H., Flores R., 3rd, Josowitz R., Krasieva T.B., Lamorte V.J., et al. UNC-51/ATG1 kinase regulates axonal transport by mediating motor-cargo assembly. Genes Dev. 2008;22:3292–3307. doi: 10.1101/gad.1734608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.McAlpine F., Williamson L.E., Tooze S.A., Chan E.Y. Regulation of nutrient-sensitive autophagy by uncoordinated 51-like kinases 1 and 2. Autophagy. 2013;9:361–373. doi: 10.4161/auto.23066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Suzuki H., Kaizuka T., Mizushima N., Noda N.N. Open and closed HORMAs regulate autophagy initiation. Autophagy. 2015;11:2123–2124. doi: 10.1080/15548627.2015.1091144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Mizushima N., Yoshimori T., Ohsumi Y. The role of Atg proteins in autophagosome formation. Annu Rev Cell Dev Biol. 2011;27:107–132. doi: 10.1146/annurev-cellbio-092910-154005. [DOI] [PubMed] [Google Scholar]
  • 27.Liang Q., Yang P., Tian E., Han J., Zhang H. The C. elegans ATG101 homolog EPG-9 directly interacts with EPG-1/Atg13 and is essential for autophagy. Autophagy. 2012;8:1426–1433. doi: 10.4161/auto.21163. [DOI] [PubMed] [Google Scholar]
  • 28.Lin L., Yang P., Huang X., Zhang H., Lu Q., Zhang H. The scaffold protein EPG-7 links cargo-receptor complexes with the autophagic assembly machinery. J Cell Biol. 2013;201:113–129. doi: 10.1083/jcb.201209098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Alers S., Löffler A.S., Paasch F., Dieterle A.M., Keppeler H., Lauber K., et al. Atg13 and FIP200 act independently of Ulk1 and Ulk2 in autophagy induction. Autophagy. 2011;7:1423–1433. doi: 10.4161/auto.7.12.18027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Wallot-Hieke N., Verma N., Schlütermann D., Berleth N., Deitersen J., Böhler P., et al. Systematic analysis of ATG13 domain requirements for autophagy induction. Autophagy. 2018;14:743–763. doi: 10.1080/15548627.2017.1387342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Hieke N., Löffler A.S., Kaizuka T., Berleth N., Böhler P., Drießen S., et al. Expression of a ULK1/2 binding-deficient ATG13 variant can partially restore autophagic activity in ATG13-deficient cells. Autophagy. 2015;11:1471–1483. doi: 10.1080/15548627.2015.1068488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kim B.W., Jin Y., Kim J., Kim J.H., Jung J., Kang S., et al. The C-terminal region of ATG101 bridges ULK1 and PtdIns3K complex in autophagy initiation. Autophagy. 2018;14:2104–2116. doi: 10.1080/15548627.2018.1504716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Hosokawa N., Sasaki T., Iemura S., Natsume T., Hara T., Mizushima N. Atg101, a novel mammalian autophagy protein interacting with Atg13. Autophagy. 2009;5:973–979. doi: 10.4161/auto.5.7.9296. [DOI] [PubMed] [Google Scholar]
  • 34.Suzuki H., Kaizuka T., Mizushima N., Noda N.N. Structure of the Atg101–Atg13 complex reveals essential roles of Atg101 in autophagy initiation. Nat Struct Mol Biol. 2015;22:572–580. doi: 10.1038/nsmb.3036. [DOI] [PubMed] [Google Scholar]
  • 35.Zhang L., Ouyang L., Guo Y., Zhang J., Liu B. UNC-51-like kinase 1: from an autophagic initiator to multifunctional drug target. J Med Chem. 2018;61:6491–6500. doi: 10.1021/acs.jmedchem.7b01684. [DOI] [PubMed] [Google Scholar]
  • 36.Popelka H., Klionsky D.J. One step closer to understanding mammalian macroautophagy initiation: interplay of 2 HORMA architectures in the ULK1 complex. Autophagy. 2015;11:1953–1955. doi: 10.1080/15548627.2015.1087635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Stjepanovic G., Davies C.W., Stanley R.E., Ragusa M.J., Kim D.J., Hurley J.H. Assembly and dynamics of the autophagy-initiating Atg1 complex. Proc Natl Acad Sci U S A. 2014;111:12793–12798. doi: 10.1073/pnas.1407214111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Liu X., Mao K., Yu A.Y.H., Omairi-Nasser A., Austin J., 2nd, Glick B.S., et al. The Atg17–Atg31–Atg29 complex coordinates with Atg11 to recruit the Vam7 SNARE and mediate autophagosome–vacuole fusion. Curr Biol. 2016;26:150–160. doi: 10.1016/j.cub.2015.11.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Tian E., Wang F., Han J., Zhang H. epg-1 functions in autophagy-regulated processes and may encode a highly divergent Atg13 homolog in C. elegans. Autophagy. 2009;5:608–615. doi: 10.4161/auto.5.5.8624. [DOI] [PubMed] [Google Scholar]
  • 40.Lai T., Garriga G. The conserved kinase UNC-51 acts with VAB-8 and UNC-14 to regulate axon outgrowth in C. elegans. Development. 2004;131:5991–6000. doi: 10.1242/dev.01457. [DOI] [PubMed] [Google Scholar]
  • 41.Su W., Liao M., Tan H., Chen Y., Zhao R., Jin W., et al. Identification of autophagic target RAB13 with small-molecule inhibitor in low-grade glioma via integrated multi-omics approaches coupled with virtual screening of traditional Chinese medicine databases. Cell Prolif. 2021;54 doi: 10.1111/cpr.13135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Pajares M., Jiménez-Moreno N., García-Yagüe Á J., Escoll M., de Ceballos M.L., Van Leuven F., et al. Transcription factor NFE2L2/NRF2 is a regulator of macroautophagy genes. Autophagy. 2016;12:1902–1916. doi: 10.1080/15548627.2016.1208889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Gothwal M., Wehrle J., Aumann K., Zimmermann V., Gründer A., Pahl H.L. A novel role for nuclear factor-erythroid 2 in erythroid maturation by modulation of mitochondrial autophagy. Haematologica. 2016;101:1054–1064. doi: 10.3324/haematol.2015.132589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Xiao Q., Liu H., Wang H.S., Cao M.T., Meng X.J., Xiang Y.L., et al. Histone deacetylase inhibitors promote epithelial–mesenchymal transition in hepatocellular carcinoma via AMPK–FOXO1–ULK1 signaling axis-mediated autophagy. Theranostics. 2020;10:10245–10261. doi: 10.7150/thno.47045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Li L., Zviti R., Ha C., Wang Z.V., Hill J.A., Lin F. Forkhead box O3 (FoxO3) regulates kidney tubular autophagy following urinary tract obstruction. J Biol Chem. 2017;292:13774–13783. doi: 10.1074/jbc.M117.791483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Lee D.H., Park S.H., Ahn J., Hong S.P., Lee E., Jang Y.J., et al. Mir214-3p and Hnf4a/Hnf4α reciprocally regulate Ulk1 expression and autophagy in nonalcoholic hepatic steatosis. Autophagy. 2021;17:2415–2431. doi: 10.1080/15548627.2020.1827779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Pike L.R., Singleton D.C., Buffa F., Abramczyk O., Phadwal K., Li J.L., et al. Transcriptional up-regulation of ULK1 by ATF4 contributes to cancer cell survival. Biochem J. 2013;449:389–400. doi: 10.1042/BJ20120972. [DOI] [PubMed] [Google Scholar]
  • 48.Goldberg A.A., Nkengfac B., Sanchez A.M.J., Moroz N., Qureshi S.T., Koromilas A.E., et al. Regulation of ULK1 expression and autophagy by STAT1. J Biol Chem. 2017;292:1899–1909. doi: 10.1074/jbc.M116.771584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Puto L.A., Brognard J., Hunter T. Transcriptional repressor DAXX promotes prostate cancer tumorigenicity via suppression of autophagy. J Biol Chem. 2015;290:15406–15420. doi: 10.1074/jbc.M115.658765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Seok S., Fu T., Choi S.E., Li Y., Zhu R., Kumar S., et al. Transcriptional regulation of autophagy by an FXR–CREB axis. Nature. 2014;516:108–111. doi: 10.1038/nature13949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Ferder I.C., Fung L., Ohguchi Y., Zhang X., Lassen K.G., Capen D., et al. Meiotic gatekeeper STRA8 suppresses autophagy by repressing Nr1d1 expression during spermatogenesis in mice. PLoS Genet. 2019;15 doi: 10.1371/journal.pgen.1008084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Ci Y., Shi K., An J., Yang Y., Hui K., Wu P., et al. ROS inhibit autophagy by downregulating ULK1 mediated by the phosphorylation of p53 in selenite-treated NB4 cells. Cell Death Dis. 2014;5:e1542. doi: 10.1038/cddis.2014.506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Conte A., Paladino S., Bianco G., Fasano D., Gerlini R., Tornincasa M., et al. High mobility group A1 protein modulates autophagy in cancer cells. Cell Death Differ. 2017;24:1948–1962. doi: 10.1038/cdd.2017.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Du Y., Guo H., Guo L., Miao J., Ren H., Liu K., et al. The regulatory effect of acetylation of HMGN2 and H3K27 on pyocyanin-induced autophagy in macrophages by affecting Ulk1 transcription. J Cell Mol Med. 2021;25:7524–7537. doi: 10.1111/jcmm.16788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Xu G., Li T., Chen J., Li C., Zhao H., Yao C., et al. Autosomal dominant retinitis pigmentosa-associated gene PRPF8 is essential for hypoxia-induced mitophagy through regulating ULK1 mRNA splicing. Autophagy. 2018;14:1818–1830. doi: 10.1080/15548627.2018.1501251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Han B., Jiang W., Cui P., Zheng K., Dang C., Wang J., et al. Microglial PGC-1α protects against ischemic brain injury by suppressing neuroinflammation. Genome Med. 2021;13:47. doi: 10.1186/s13073-021-00863-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Charbe N.B., Amnerkar N.D., Ramesh B., Tambuwala M.M., Bakshi H.A., Aljabali A.A.A., et al. Small interfering RNA for cancer treatment: overcoming hurdles in delivery. Acta Pharm Sin B. 2020;10:2075–2109. doi: 10.1016/j.apsb.2020.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Lahiri V., Metur S.P., Hu Z., Song X., Mari M., Hawkins W.D., et al. Post-transcriptional regulation of ATG1 is a critical node that modulates autophagy during distinct nutrient stresses. Autophagy. 2022;18:1694–1714. doi: 10.1080/15548627.2021.1997305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Zou J., Zhu X., Xiang D., Zhang Y., Li J., Su Z., et al. LIX1-like protein promotes liver cancer progression via miR-21-3p-mediated inhibition of fructose-1,6-bisphosphatase. Acta Pharm Sin B. 2021;11:1578–1591. doi: 10.1016/j.apsb.2021.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Li X., Tian Y., Tu M.J., Ho P.Y., Batra N., Yu A.M. Bioengineered miR-27b-3p and miR-328-3p modulate drug metabolism and disposition via the regulation of target ADME gene expression. Acta Pharm Sin B. 2019;9:639–647. doi: 10.1016/j.apsb.2018.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Chen H., Zhang Z., Lu Y., Song K., Liu X., Xia F., et al. Downregulation of ULK1 by microRNA-372 inhibits the survival of human pancreatic adenocarcinoma cells. Cancer Sci. 2017;108:1811–1819. doi: 10.1111/cas.13315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Li W., Yang Y., Ba Z., Li S., Chen H., Hou X., et al. MicroRNA-93 regulates hypoxia-induced autophagy by targeting ULK1. Oxid Med Cell Longev. 2017;2017 doi: 10.1155/2017/2709053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Li Z., Lan Y., Zhao K., Lv X., Ding N., Lu H., et al. miR-142-5p disrupts neuronal morphogenesis underlying porcine hemagglutinating encephalomyelitis virus infection by targeting Ulk1. Front Cell Infect Microbiol. 2017;7:155. doi: 10.3389/fcimb.2017.00155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Füllgrabe J., Klionsky D.J., Joseph B. The return of the nucleus: transcriptional and epigenetic control of autophagy. Nat Rev Mol Cell Biol. 2014;15:65–74. doi: 10.1038/nrm3716. [DOI] [PubMed] [Google Scholar]
  • 65.Wu H., Wang F., Hu S., Yin C., Li X., Zhao S., et al. MiR-20a and miR-106b negatively regulate autophagy induced by leucine deprivation via suppression of ULK1 expression in C2C12 myoblasts. Cell Signal. 2012;24:2179–2186. doi: 10.1016/j.cellsig.2012.07.001. [DOI] [PubMed] [Google Scholar]
  • 66.Rothschild S.I., Gautschi O., Batliner J., Gugger M., Fey M.F., Tschan M.P. MicroRNA-106a targets autophagy and enhances sensitivity of lung cancer cells to Src inhibitors. Lung Cancer. 2017;107:73–83. doi: 10.1016/j.lungcan.2016.06.004. [DOI] [PubMed] [Google Scholar]
  • 67.Liu K., Hong D., Zhang F., Li X., He M., Han X., et al. MicroRNA-106a inhibits autophagy process and antimicrobial responses by targeting ULK1, ATG7, and ATG16L1 during mycobacterial infection. Front Immunol. 2020;11 doi: 10.3389/fimmu.2020.610021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Salgado-García R., Coronel-Hernández J., Delgado-Waldo I., Cantú de León D., García-Castillo V., López-Urrutia E., et al. Negative regulation of ULK1 by microRNA-106a in autophagy induced by a triple drug combination in colorectal cancer cells in vitro. Genes (Basel) 2021;12:245. doi: 10.3390/genes12020245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Jin F., Wang Y., Li M., Zhu Y., Liang H., Wang C., et al. MiR-26 enhances chemosensitivity and promotes apoptosis of hepatocellular carcinoma cells through inhibiting autophagy. Cell Death Dis. 2017;8 doi: 10.1038/cddis.2016.461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Soni M., Patel Y., Markoutsa E., Jie C., Liu S., Xu P., et al. Autophagy, cell viability, and chemoresistance are regulated by miR-489 in breast cancer. Mol Cancer Res. 2018;16:1348–1360. doi: 10.1158/1541-7786.MCR-17-0634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Xie K., Chen M., Zhu M., Wang C., Qin N., Liang C., et al. A polymorphism in miR-1262 regulatory region confers the risk of lung cancer in Chinese population. Int J Cancer. 2017;141:958–966. doi: 10.1002/ijc.30788. [DOI] [PubMed] [Google Scholar]
  • 72.Ma Z., Li L., Livingston M.J., Zhang D., Mi Q., Zhang M., et al. p53/microRNA-214/ULK1 axis impairs renal tubular autophagy in diabetic kidney disease. J Clin Invest. 2020;130:5011–5026. doi: 10.1172/JCI135536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.D'Adamo S., Alvarez-Garcia O., Muramatsu Y., Flamigni F., Lotz M.K. MicroRNA-155 suppresses autophagy in chondrocytes by modulating expression of autophagy proteins. Osteoarthritis Cartilage. 2016;24:1082–1091. doi: 10.1016/j.joca.2016.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Huang F., Chen W., Peng J., Li Y., Zhuang Y., Zhu Z., et al. LncRNA PVT1 triggers Cyto-protective autophagy and promotes pancreatic ductal adenocarcinoma development via the miR-20a-5p/ULK1 axis. Mol Cancer. 2018;17:98. doi: 10.1186/s12943-018-0845-6. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 75.Guo X., Wu X., Han Y., Tian E., Cheng J. LncRNA MALAT1 protects cardiomyocytes from isoproterenol-induced apoptosis through sponging miR-558 to enhance ULK1-mediated protective autophagy. J Cell Physiol. 2019;234:10842–10854. doi: 10.1002/jcp.27925. [DOI] [PubMed] [Google Scholar]
  • 76.Wang X., Lan Z., He J., Lai Q., Yao X., Li Q., et al. LncRNA SNHG6 promotes chemoresistance through ULK1-induced autophagy by sponging miR-26a-5p in colorectal cancer cells. Cancer Cell Int. 2019;19:234. doi: 10.1186/s12935-019-0951-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Zheng L., Lin S., Lv C. MiR-26a-5p regulates cardiac fibroblasts collagen expression by targeting ULK1. Sci Rep. 2018;8:2104. doi: 10.1038/s41598-018-20561-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Zhu X., Yang G., Xu J., Zhang C. Silencing of SNHG6 induced cell autophagy by targeting miR-26a-5p/ULK1 signaling pathway in human osteosarcoma. Cancer Cell Int. 2019;19:82. doi: 10.1186/s12935-019-0794-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Song C., Qi H., Liu Y., Chen Y., Shi P., Zhang S., et al. Inhibition of lncRNA Gm15834 attenuates autophagy-mediated myocardial hypertrophy via the miR-30b-3p/ULK1 axis in mice. Mol Ther. 2021;29:1120–1137. doi: 10.1016/j.ymthe.2020.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Jiang F., Lou J., Zheng X.M., Yang X.Y. LncRNA MIAT regulates autophagy and apoptosis of macrophage infected by Mycobacterium tuberculosis through the miR-665/ULK1 signaling axis. Mol Immunol. 2021;139:42–49. doi: 10.1016/j.molimm.2021.07.023. [DOI] [PubMed] [Google Scholar]
  • 81.Chen C.L., Tseng Y.W., Wu J.C., Chen G.Y., Lin K.C., Hwang S.M., et al. Suppression of hepatocellular carcinoma by baculovirus-mediated expression of long non-coding RNA PTENP1 and microRNA regulation. Biomaterials. 2015;44:71–81. doi: 10.1016/j.biomaterials.2014.12.023. [DOI] [PubMed] [Google Scholar]
  • 82.Zhu K., Yuan Y., Wen J., Chen D., Zhu W., Ouyang Z., et al. LncRNA Sox2OT-V7 promotes doxorubicin-induced autophagy and chemoresistance in osteosarcoma via tumor-suppressive miR-142/miR-22. Aging (Albany N Y) 2020;12:6644–6666. doi: 10.18632/aging.103004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Cao H.X., Miao C.F., Sang L.N., Huang Y.M., Zhang R., Sun L., et al. Circ_0009910 promotes imatinib resistance through ULK1-induced autophagy by sponging miR-34a-5p in chronic myeloid leukemia. Life Sci. 2020;243 doi: 10.1016/j.lfs.2020.117255. [DOI] [PubMed] [Google Scholar]
  • 84.Liang G., Ling Y., Mehrpour M., Saw P.E., Liu Z., Tan W., et al. Autophagy-associated circRNA circCDYL augments autophagy and promotes breast cancer progression. Mol Cancer. 2020;19:65. doi: 10.1186/s12943-020-01152-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Shen Q., Xu Z., Xu S. Long non-coding RNA LUCAT1 contributes to cisplatin resistance by regulating the miR-514a-3p/ULK1 axis in human non-small cell lung cancer. Int J Oncol. 2020;57:967–979. doi: 10.3892/ijo.2020.5106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Zhao J., Yang M., Li Q., Pei X., Zhu X. miR-132-5p regulates apoptosis and autophagy in MPTP model of Parkinson's disease by targeting ULK1. Neuroreport. 2020;31:959–965. doi: 10.1097/WNR.0000000000001494. [DOI] [PubMed] [Google Scholar]
  • 87.Du L., Liu W., Aldana-Masangkay G., Pozhitkov A., Pichiorri F., Chen Y., et al. SUMOylation inhibition enhances dexamethasone sensitivity in multiple myeloma. J Exp Clin Cancer Res. 2022;41:8. doi: 10.1186/s13046-021-02226-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Jin J., Britschgi A., Schläfli A.M., Humbert M., Shan-Krauer D., Batliner J., et al. Low autophagy (ATG) gene expression is associated with an immature AML blast cell phenotype and can be restored during AML differentiation therapy. Oxid Med Cell Longev. 2018;2018 doi: 10.1155/2018/1482795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Wang H., Wang X., Zhang H., Deng T., Liu R., Liu Y., et al. The HSF1/miR-135b-5p axis induces protective autophagy to promote oxaliplatin resistance through the MUL1/ULK1 pathway in colorectal cancer. Oncogene. 2021;40:4695–4708. doi: 10.1038/s41388-021-01898-z. [DOI] [PubMed] [Google Scholar]
  • 90.Li Y., Jiang J., Liu W., Wang H., Zhao L., Liu S., et al. MicroRNA-378 promotes autophagy and inhibits apoptosis in skeletal muscle. Proc Natl Acad Sci U S A. 2018;115 doi: 10.1073/pnas.1803377115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Tomasetti M., Monaco F., Manzella N., Rohlena J., Rohlenova K., Staffolani S., et al. MicroRNA-126 induces autophagy by altering cell metabolism in malignant mesothelioma. Oncotarget. 2016;7:36338–36352. doi: 10.18632/oncotarget.8916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Li S., Zeng X., Ma R., Wang L. MicroRNA-21 promotes the proliferation, migration and invasion of non-small cell lung cancer A549 cells by regulating autophagy activity via AMPK/ULK1 signaling pathway. Exp Ther Med. 2018;16:2038–2045. doi: 10.3892/etm.2018.6370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Lin Y., Deng W., Pang J., Kemper T., Hu J., Yin J., et al. The microRNA-99 family modulates hepatitis B virus replication by promoting IGF-1R/PI3K/Akt/mTOR/ULK1 signaling-induced autophagy. Cell Microbiol. 2017;19 doi: 10.1111/cmi.12709. [DOI] [PubMed] [Google Scholar]
  • 94.Lv Q., Zhong Z., Hu B., Yan S., Yan Y., Zhang J., et al. MicroRNA-3473b regulates the expression of TREM2/ULK1 and inhibits autophagy in inflammatory pathogenesis of Parkinson disease. J Neurochem. 2021;157:599–610. doi: 10.1111/jnc.15299. [DOI] [PubMed] [Google Scholar]
  • 95.Wang Y., Fang Z., Hong M., Yang D., Xie W. Long-noncoding RNAs (lncRNAs) in drug metabolism and disposition, implications in cancer chemo-resistance. Acta Pharm Sin B. 2020;10:105–112. doi: 10.1016/j.apsb.2019.09.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Chen Z.H., Wang W.T., Huang W., Fang K., Sun Y.M., Liu S.R., et al. The lncRNA HOTAIRM1 regulates the degradation of PML-RARA oncoprotein and myeloid cell differentiation by enhancing the autophagy pathway. Cell Death Differ. 2017;24:212–224. doi: 10.1038/cdd.2016.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Liu F., Ai F.Y., Zhang D.C., Tian L., Yang Z.Y., Liu S.J. LncRNA NEAT1 knockdown attenuates autophagy to elevate 5-FU sensitivity in colorectal cancer via targeting miR-34a. Cancer Med. 2020;9:1079–1091. doi: 10.1002/cam4.2746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Li G., Qian L., Tang X., Chen Y., Zhao Z., Zhang C. Long non-coding RNA growth arrest-specific 5 (GAS5) acts as a tumor suppressor by promoting autophagy in breast cancer. Mol Med Rep. 2020;22:2460–2468. doi: 10.3892/mmr.2020.11334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Zhan S., Wang K., Xiang Q., Song Y., Li S., Liang H., et al. lncRNA HOTAIR upregulates autophagy to promote apoptosis and senescence of nucleus pulposus cells. J Cell Physiol. 2020;235:2195–2208. doi: 10.1002/jcp.29129. [DOI] [PubMed] [Google Scholar]
  • 100.Han S., Li X., Wang K., Zhu D., Meng B., Liu J., et al. PURPL represses autophagic cell death to promote cutaneous melanoma by modulating ULK1 phosphorylation. Cell Death Dis. 2021;12:1070. doi: 10.1038/s41419-021-04362-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Liu J., Ren L., Li S., Li W., Zheng X., Yang Y., et al. The biology, function, and applications of exosomes in cancer. Acta Pharm Sin B. 2021;11:2783–2797. doi: 10.1016/j.apsb.2021.01.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Wang H., Sun G., Xu P., Lv J., Zhang X., Zhang L., et al. Circular RNA TMEM87A promotes cell proliferation and metastasis of gastric cancer by elevating ULK1 via sponging miR-142-5p. J Gastroenterol. 2021;56:125–138. doi: 10.1007/s00535-020-01744-1. [DOI] [PubMed] [Google Scholar]
  • 103.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. doi: 10.1038/ncb2152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Yao F., Zhang M., Chen L. 5′-Monophosphate-activated protein kinase (AMPK) improves autophagic activity in diabetes and diabetic complications. Acta Pharm Sin B. 2016;6:20–25. doi: 10.1016/j.apsb.2015.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Shang L., Chen S., Du F., Li S., Zhao L., Wang X. Nutrient starvation elicits an acute autophagic response mediated by Ulk1 dephosphorylation and its subsequent dissociation from AMPK. Proc Natl Acad Sci U S A. 2011;108:4788–4793. doi: 10.1073/pnas.1100844108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Gong J., Gu H., Zhao L., Wang L., Liu P., Wang F., et al. Phosphorylation of ULK1 by AMPK is essential for mouse embryonic stem cell self-renewal and pluripotency. Cell Death Dis. 2018;9:38. doi: 10.1038/s41419-017-0054-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Dunlop E.A., Hunt D.K., Acosta-Jaquez H.A., Fingar D.C., Tee A.R. ULK1 inhibits mTORC1 signaling, promotes multisite Raptor phosphorylation and hinders substrate binding. Autophagy. 2011;7:737–747. doi: 10.4161/auto.7.7.15491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Colecchia D., Dapporto F., Tronnolone S., Salvini L., Chiariello M. MAPK15 is part of the ULK complex and controls its activity to regulate early phases of the autophagic process. J Biol Chem. 2018;293:15962–15976. doi: 10.1074/jbc.RA118.002527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Zhang M., Fang L., Zhou L., Molino A., Valentino M.R., Yang S., et al. MAPK15–ULK1 signaling regulates mitophagy of airway epithelial cell in chronic obstructive pulmonary disease. Free Radic Biol Med. 2021;172:541–549. doi: 10.1016/j.freeradbiomed.2021.07.004. [DOI] [PubMed] [Google Scholar]
  • 110.Ryu H.Y., Kim L.E., Jeong H., Yeo B.K., Lee J.W., Nam H., et al. GSK3B induces autophagy by phosphorylating ULK1. Exp Mol Med. 2021;53:369–383. doi: 10.1038/s12276-021-00570-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Wang C., Wang H., Zhang D., Luo W., Liu R., Xu D., et al. Phosphorylation of ULK1 affects autophagosome fusion and links chaperone-mediated autophagy to macroautophagy. Nat Commun. 2018;9:3492. doi: 10.1038/s41467-018-05449-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Miao H., Qiu F., Huang B., Liu X., Zhang H., Liu Z., et al. PKCα replaces AMPK to regulate mitophagy: another PEDF role on ischaemic cardioprotection. J Cell Mol Med. 2018;22:5732–5742. doi: 10.1111/jcmm.13849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Gallolu Kankanamalage S., Lee A.Y., Wichaidit C., Lorente-Rodriguez A., Shah A.M., Stippec S., et al. WNK1 is an unexpected autophagy inhibitor. Autophagy. 2017;13:969–970. doi: 10.1080/15548627.2017.1286431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Gallolu Kankanamalage S., Lee A.Y., Wichaidit C., Lorente-Rodriguez A., Shah A.M., Stippec S., et al. Multistep regulation of autophagy by WNK1. Proc Natl Acad Sci U S A. 2016;113:14342–14347. doi: 10.1073/pnas.1617649113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Hu Z., Sankar D.S., Vu B., Leytens A., Vionnet C., Wu W., et al. ULK1 phosphorylation of striatin activates protein phosphatase 2A and autophagy. Cell Rep. 2021;36 doi: 10.1016/j.celrep.2021.109762. [DOI] [PubMed] [Google Scholar]
  • 116.Kim J.H., Seo D., Kim S.J., Choi D.W., Park J.S., Ha J., et al. The deubiquitinating enzyme USP20 stabilizes ULK1 and promotes autophagy initiation. EMBO Rep. 2018;19 doi: 10.15252/embr.201744378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Thayer J.A., Awad O., Hegdekar N., Sarkar C., Tesfay H., Burt C., et al. The PARK10 gene USP24 is a negative regulator of autophagy and ULK1 protein stability. Autophagy. 2020;16:140–153. doi: 10.1080/15548627.2019.1598754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Raimondi M., Cesselli D., Di Loreto C., La Marra F., Schneider C., Demarchi F. USP1 (ubiquitin specific peptidase 1) targets ULK1 and regulates its cellular compartmentalization and autophagy. Autophagy. 2019;15:613–630. doi: 10.1080/15548627.2018.1535291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Nazio F., Carinci M., Valacca C., Bielli P., Strappazzon F., Antonioli M., et al. Fine-tuning of ULK1 mRNA and protein levels is required for autophagy oscillation. J Cell Biol. 2016;215:841–856. doi: 10.1083/jcb.201605089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Han S.H., Korm S., Han Y.G., Choi S.Y., Kim S.H., Chung H.J., et al. GCA links TRAF6–ULK1-dependent autophagy activation in resistant chronic myeloid leukemia. Autophagy. 2019;15:2076–2090. doi: 10.1080/15548627.2019.1596492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Nazio F., Strappazzon F., Antonioli M., Bielli P., Cianfanelli V., Bordi M., et al. mTOR inhibits autophagy by controlling ULK1 ubiquitylation, self-association and function through AMBRA1 and TRAF6. Nat Cell Biol. 2013;15:406–416. doi: 10.1038/ncb2708. [DOI] [PubMed] [Google Scholar]
  • 122.Di Rienzo M., Piacentini M., Fimia G.M. A TRIM32–AMBRA1–ULK1 complex initiates the autophagy response in atrophic muscle cells. Autophagy. 2019;15:1674–1676. doi: 10.1080/15548627.2019.1635385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Tian S., Jin S., Wu Y., Liu T., Luo M., Ou J., et al. High-throughput screening of functional deubiquitinating enzymes in autophagy. Autophagy. 2021;17:1367–1378. doi: 10.1080/15548627.2020.1761652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Zhou X., Babu J.R., da Silva S., Shu Q., Graef I.A., Oliver T., et al. Unc-51-like kinase 1/2-mediated endocytic processes regulate filopodia extension and branching of sensory axons. Proc Natl Acad Sci U S A. 2007;104:5842–5847. doi: 10.1073/pnas.0701402104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Chauhan S., Kumar S., Jain A., Ponpuak M., Mudd M.H., Kimura T., et al. TRIMs and Galectins globally cooperate and TRIM16 and Galectin-3 co-direct autophagy in endomembrane damage homeostasis. Dev Cell. 2016;39:13–27. doi: 10.1016/j.devcel.2016.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Li J., Zhang T., Ren T., Liao X., Hao Y., Lim J.S., et al. Oxygen-sensitive methylation of ULK1 is required for hypoxia-induced autophagy. Nat Commun. 2022;13:1172. doi: 10.1038/s41467-022-28831-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Lin S.Y., Li T.Y., Liu Q., Zhang C., Li X., Chen Y., et al. GSK3–TIP60–ULK1 signaling pathway links growth factor deprivation to autophagy. Science. 2012;336:477–481. doi: 10.1126/science.1217032. [DOI] [PubMed] [Google Scholar]
  • 128.Wang Y., Chen Q., Jiao F., Shi C., Pei M., Wang L., et al. Histone deacetylase 2 regulates ULK1 mediated pyroptosis during acute liver failure by the K68 acetylation site. Cell Death Dis. 2021;12:55. doi: 10.1038/s41419-020-03317-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Pyo K.E., Kim C.R., Lee M., Kim J.S., Kim K.I., Baek S.H. ULK1 O-GlcNAcylation is crucial for activating VPS34 via ATG14L during autophagy initiation. Cell Rep. 2018;25:2878–2890.e4. doi: 10.1016/j.celrep.2018.11.042. [DOI] [PubMed] [Google Scholar]
  • 130.Nguyen T.T.P., Kim D.Y., Lee Y.G., Lee Y.S., Truong X.T., Lee J.H., et al. SREBP-1c impairs ULK1 sulfhydration-mediated autophagic flux to promote hepatic steatosis in high-fat-diet-fed mice. Mol Cell. 2021;81:3820–3832.e7. doi: 10.1016/j.molcel.2021.06.003. [DOI] [PubMed] [Google Scholar]
  • 131.Wold M.S., Lim J., Lachance V., Deng Z., Yue Z. ULK1-mediated phosphorylation of ATG14 promotes autophagy and is impaired in Huntington's disease models. Mol Neurodegener. 2016;11:76. doi: 10.1186/s13024-016-0141-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Park J.M., Seo M., Jung C.H., Grunwald D., Stone M., Otto N.M., et al. ULK1 phosphorylates Ser30 of BECN1 in association with ATG14 to stimulate autophagy induction. Autophagy. 2018;14:584–597. doi: 10.1080/15548627.2017.1422851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Mercer T.J., Ohashi Y., Boeing S., Jefferies H.B.J., De Tito S., Flynn H., et al. Phosphoproteomic identification of ULK substrates reveals VPS15-dependent ULK/VPS34 interplay in the regulation of autophagy. EMBO J. 2021;40 doi: 10.15252/embj.2020105985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Zhou C., Ma K., Gao R., Mu C., Chen L., Liu Q., et al. Regulation of mATG9 trafficking by Src- and ULK1-mediated phosphorylation in basal and starvation-induced autophagy. Cell Res. 2017;27:184–201. doi: 10.1038/cr.2016.146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Xu J., Fotouhi M., McPherson P.S. Phosphorylation of the exchange factor DENND3 by ULK in response to starvation activates Rab12 and induces autophagy. EMBO Rep. 2015;16:709–718. doi: 10.15252/embr.201440006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Yang Y., Klionsky D.J. An AMPK–ULK1–PIKFYVE signaling axis for PtdIns5P-dependent autophagy regulation upon glucose starvation. Autophagy. 2021;17:2663–2664. doi: 10.1080/15548627.2021.1959240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Karabiyik C., Rubinsztein D.C. AMPK-activated ULK1 phosphorylates PIKFYVE to drive formation of PtdIns5P-containing autophagosomes during glucose starvation. Autophagy. 2021;17:3877–3878. doi: 10.1080/15548627.2021.1961409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Karabiyik C., Vicinanza M., Son S.M., Rubinsztein D.C. Glucose starvation induces autophagy via ULK1-mediated activation of PIKfyve in an AMPK-dependent manner. Dev Cell. 2021;56:1961–1975.e5. doi: 10.1016/j.devcel.2021.05.010. [DOI] [PubMed] [Google Scholar]
  • 139.Liu C.C., Lin Y.C., Chen Y.H., Chen C.M., Pang L.Y., Chen H.A., et al. Cul3-KLHL20 ubiquitin ligase governs the turnover of ULK1 and VPS34 complexes to control autophagy termination. Mol Cell. 2016;61:84–97. doi: 10.1016/j.molcel.2015.11.001. [DOI] [PubMed] [Google Scholar]
  • 140.Jeong Y.T., Simoneschi D., Keegan S., Melville D., Adler N.S., Saraf A., et al. The ULK1–FBXW5–SEC23B nexus controls autophagy. Elife. 2018;7 doi: 10.7554/eLife.42253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Di Bartolomeo S., Corazzari M., Nazio F., Oliverio S., Lisi G., Antonioli M., et al. The dynamic interaction of AMBRA1 with the dynein motor complex regulates mammalian autophagy. J Cell Biol. 2010;191:155–168. doi: 10.1083/jcb.201002100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Pengo N., Agrotis A., Prak K., Jones J., Ketteler R. A reversible phospho-switch mediated by ULK1 regulates the activity of autophagy protease ATG4B. Nat Commun. 2017;8:294. doi: 10.1038/s41467-017-00303-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Dunlop E.A., Seifan S., Claessens T., Behrends C., Kamps M.A., Rozycka E., et al. FLCN, a novel autophagy component, interacts with GABARAP and is regulated by ULK1 phosphorylation. Autophagy. 2014;10:1749–1760. doi: 10.4161/auto.29640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Siva Sankar D., Hu Z., Dengjel J. The complex interplay between ULK1 and protein phosphatases in autophagy regulation. Autophagy. 2022;18:455–456. doi: 10.1080/15548627.2021.2002546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Lee D.H., Park J.S., Lee Y.S., Han J., Lee D.K., Kwon S.W., et al. SQSTM1/p62 activates NFE2L2/NRF2 via ULK1-mediated autophagic KEAP1 degradation and protects mouse liver from lipotoxicity. Autophagy. 2020;16:1949–1973. doi: 10.1080/15548627.2020.1712108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Lee D.H., Park J.S., Lee Y.S., Bae S.H. PERK prevents hepatic lipotoxicity by activating the p62–ULK1 axis-mediated noncanonical KEAP1–Nrf2 pathway. Redox Biol. 2022;50 doi: 10.1016/j.redox.2022.102235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Torii S., Yoshida T., Arakawa S., Honda S., Nakanishi A., Shimizu S. Identification of PPM1D as an essential Ulk1 phosphatase for genotoxic stress-induced autophagy. EMBO Rep. 2016;17:1552–1564. doi: 10.15252/embr.201642565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Li G.M., Li L., Li M.Q., Chen X., Su Q., Deng Z.J., et al. DAPK3 inhibits gastric cancer progression via activation of ULK1-dependent autophagy. Cell Death Differ. 2021;28:952–967. doi: 10.1038/s41418-020-00627-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Lu H., Xiao J., Ke C., Ni X., Xiu R., Tian Q., et al. TOPK inhibits autophagy by phosphorylating ULK1 and promotes glioma resistance to TMZ. Cell Death Dis. 2019;10:583. doi: 10.1038/s41419-019-1805-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.He Y., She H., Zhang T., Xu H., Cheng L., Yepes M., et al. p38 MAPK inhibits autophagy and promotes microglial inflammatory responses by phosphorylating ULK1. J Cell Biol. 2018;217:315–328. doi: 10.1083/jcb.201701049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Man N., Tan Y., Sun X.J., Liu F., Cheng G., Greenblatt S.M., et al. Caspase-3 controls AML1–ETO-driven leukemogenesis via autophagy modulation in a ULK1-dependent manner. Blood. 2017;129:2782–2792. doi: 10.1182/blood-2016-10-745034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Lim J., Lachenmayer M.L., Wu S., Liu W., Kundu M., Wang R., et al. Proteotoxic stress induces phosphorylation of p62/SQSTM1 by ULK1 to regulate selective autophagic clearance of protein aggregates. PLoS Genet. 2015;11 doi: 10.1371/journal.pgen.1004987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Wu M.Y., Wang E.J., Feng D., Li M., Ye R.D., Lu J.H. Pharmacological insights into autophagy modulation in autoimmune diseases. Acta Pharm Sin B. 2021;11:3364–3378. doi: 10.1016/j.apsb.2021.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Joo J.H., Wang B., Frankel E., Ge L., Xu L., Iyengar R., et al. The noncanonical role of ULK/ATG1 in ER-to-Golgi trafficking is essential for cellular homeostasis. Mol Cell. 2016;62:491–506. doi: 10.1016/j.molcel.2016.04.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Li J., Qi W., Chen G., Feng D., Liu J., Ma B., et al. Mitochondrial outer-membrane E3 ligase MUL1 ubiquitinates ULK1 and regulates selenite-induced mitophagy. Autophagy. 2015;11:1216–1229. doi: 10.1080/15548627.2015.1017180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Deng R., Zhang H.L., Huang J.H., Cai R.Z., Wang Y., Chen Y.H., et al. MAPK1/3 kinase-dependent ULK1 degradation attenuates mitophagy and promotes breast cancer bone metastasis. Autophagy. 2021;17:3011–3029. doi: 10.1080/15548627.2020.1850609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Lee J.H., Oh S.J., Yun J., Shin O.S. Nonstructural protein NS1 of influenza virus disrupts mitochondrial dynamics and enhances mitophagy via ULK1 and BNIP3. Viruses. 2021;13:1845. doi: 10.3390/v13091845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Wu W., Tian W., Hu Z., Chen G., Huang L., Li W., et al. ULK1 translocates to mitochondria and phosphorylates FUNDC1 to regulate mitophagy. EMBO Rep. 2014;15:566–575. doi: 10.1002/embr.201438501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Kumar A., Shaha C. SESN2 facilitates mitophagy by helping Parkin translocation through ULK1 mediated Beclin1 phosphorylation. Sci Rep. 2018;8:615. doi: 10.1038/s41598-017-19102-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Saito T., Nah J., Oka S.I., Mukai R., Monden Y., Maejima Y., et al. An alternative mitophagy pathway mediated by Rab9 protects the heart against ischemia. J Clin Invest. 2019;129:802–819. doi: 10.1172/JCI122035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.He J.P., Hou P.P., Chen Q.T., Wang W.J., Sun X.Y., Yang P.B., et al. Flightless-I blocks p62-mediated recognition of LC3 to impede selective autophagy and promote breast cancer progression. Cancer Res. 2018;78:4853–4864. doi: 10.1158/0008-5472.CAN-17-3835. [DOI] [PubMed] [Google Scholar]
  • 162.Ro S.H., Semple I.A., Park H., Park H., Park H.W., Kim M., et al. Sestrin2 promotes Unc-51-like kinase 1 mediated phosphorylation of p62/sequestosome-1. FEBS J. 2014;281:3816–3827. doi: 10.1111/febs.12905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Tang Y., Wang L., Yi T., Xu J., Wang J., Qin J.J., et al. Synergistic effects of autophagy/mitophagy inhibitors and magnolol promote apoptosis and antitumor efficacy. Acta Pharm Sin B. 2021;11:3966–3982. doi: 10.1016/j.apsb.2021.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Zhao P., Wong K.I., Sun X., Reilly S.M., Uhm M., Liao Z., et al. TBK1 at the crossroads of inflammation and energy homeostasis in adipose tissue. Cell. 2018;172:731–743.e12. doi: 10.1016/j.cell.2018.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Bach M., Larance M., James D.E., Ramm G. The serine/threonine kinase ULK1 is a target of multiple phosphorylation events. Biochem J. 2011;440:283–291. doi: 10.1042/BJ20101894. [DOI] [PubMed] [Google Scholar]
  • 166.Shen Y., Liu W.W., Zhang X., Shi J.G., Jiang S., Zheng L., et al. TRAF3 promotes ROS production and pyroptosis by targeting ULK1 ubiquitination in macrophages. FASEB J. 2020;34:7144–7159. doi: 10.1096/fj.201903073R. [DOI] [PubMed] [Google Scholar]
  • 167.Slobodnyuk K., Radic N., Ivanova S., Llado A., Trempolec N., Zorzano A., et al. Autophagy-induced senescence is regulated by p38α signaling. Cell Death Dis. 2019;10:376. doi: 10.1038/s41419-019-1607-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Li Z., Tian X., Ji X., Wang J., Chen H., Wang D., et al. ULK1–ATG13 and their mitotic phospho-regulation by CDK1 connect autophagy to cell cycle. PLoS Biol. 2020;18 doi: 10.1371/journal.pbio.3000288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Zhang C.S., Hardie D.G., Lin S.C. Glucose starvation blocks translation at multiple levels. Cell Metab. 2020;31:217–218. doi: 10.1016/j.cmet.2020.01.005. [DOI] [PubMed] [Google Scholar]
  • 170.Yoon I., Nam M., Kim H.K., Moon H.S., Kim S., Jang J., et al. Glucose-dependent control of leucine metabolism by leucyl-tRNA synthetase 1. Science. 2020;367:205–210. doi: 10.1126/science.aau2753. [DOI] [PubMed] [Google Scholar]
  • 171.Li T.Y., Sun Y., Liang Y., Liu Q., Shi Y., Zhang C.S., et al. ULK1/2 constitute a bifurcate node controlling glucose metabolic fluxes in addition to autophagy. Mol Cell. 2016;62:359–370. doi: 10.1016/j.molcel.2016.04.009. [DOI] [PubMed] [Google Scholar]
  • 172.Nusse R., Clevers H. Wnt/β-catenin signaling, disease, and emerging therapeutic modalities. Cell. 2017;169:985–999. doi: 10.1016/j.cell.2017.05.016. [DOI] [PubMed] [Google Scholar]
  • 173.Hwang S.H., Bang S., Kang K.S., Kang D., Chung J. ULK1 negatively regulates Wnt signaling by phosphorylating dishevelled. Biochem Biophys Res Commun. 2019;508:308–313. doi: 10.1016/j.bbrc.2018.11.139. [DOI] [PubMed] [Google Scholar]
  • 174.Mao L., Zhan Y.Y., Wu B., Yu Q., Xu L., Hong X., et al. ULK1 phosphorylates Exo70 to suppress breast cancer metastasis. Nat Commun. 2020;11:117. doi: 10.1038/s41467-019-13923-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Li R., Yuan F., Fu W., Zhang L., Zhang N., Wang Y., et al. Serine/threonine kinase Unc-51-like kinase-1 (Ulk1) phosphorylates the co-chaperone cell division cycle protein 37 (Cdc37) and thereby disrupts the stability of Cdc37 client proteins. J Biol Chem. 2017;292:2830–2841. doi: 10.1074/jbc.M116.762443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Yuan F., Jin X., Li D., Song Y., Zhang N., Yang X., et al. ULK1 phosphorylates Mad1 to regulate spindle assembly checkpoint. Nucleic Acids Res. 2019;47:8096–8110. doi: 10.1093/nar/gkz602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Wang B., Maxwell B.A., Joo J.H., Gwon Y., Messing J., Mishra A., et al. ULK1 and ULK2 regulate stress granule disassembly through phosphorylation and activation of VCP/p97. Mol Cell. 2019;74:742–757.e8. doi: 10.1016/j.molcel.2019.03.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Konno H., Konno K., Barber G.N. Cyclic dinucleotides trigger ULK1 (ATG1) phosphorylation of STING to prevent sustained innate immune signaling. Cell. 2013;155:688–698. doi: 10.1016/j.cell.2013.09.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Rickard J.A., O'Donnell J.A., Evans J.M., Lalaoui N., Poh A.R., Rogers T., et al. RIPK1 regulates RIPK3–MLKL-driven systemic inflammation and emergency hematopoiesis. Cell. 2014;157:1175–1188. doi: 10.1016/j.cell.2014.04.019. [DOI] [PubMed] [Google Scholar]
  • 180.Wu W., Wang X., Berleth N., Deitersen J., Wallot-Hieke N., Böhler P., et al. The autophagy-initiating kinase ULK1 controls RIPK1-mediated cell death. Cell Rep. 2020;31 doi: 10.1016/j.celrep.2020.107547. [DOI] [PubMed] [Google Scholar]
  • 181.Tsang T., Posimo J.M., Gudiel A.A., Cicchini M., Feldser D.M., Brady D.C. Copper is an essential regulator of the autophagic kinases ULK1/2 to drive lung adenocarcinoma. Nat Cell Biol. 2020;22:412–424. doi: 10.1038/s41556-020-0481-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Grunwald D.S., Otto N.M., Park J.M., Song D., Kim D.H. GABARAPs and LC3s have opposite roles in regulating ULK1 for autophagy induction. Autophagy. 2020;16:600–614. doi: 10.1080/15548627.2019.1632620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Carinci M., Testa B., Bordi M., Milletti G., Bonora M., Antonucci L., et al. TFG binds LC3C to regulate ULK1 localization and autophagosome formation. EMBO J. 2021;40 doi: 10.15252/embj.2019103563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Chen Y.D., Fang Y.T., Chang C.P., Lin C.F., Hsu L.J., Wu S.R., et al. S100A10 regulates ULK1 localization to ER–mitochondria contact sites in IFN-γ-triggered autophagy. J Mol Biol. 2017;429:142–157. doi: 10.1016/j.jmb.2016.11.009. [DOI] [PubMed] [Google Scholar]
  • 185.Jia Y., Li H.Y., Wang Y., Wang J., Zhu J.W., Wei Y.Y., et al. Crosstalk between hypoxia-sensing ULK1/2 and YAP-driven glycolysis fuels pancreatic ductal adenocarcinoma development. Int J Biol Sci. 2021;17:2772–2794. doi: 10.7150/ijbs.60018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Ho W.Y., Tai Y.K., Chang J.C., Liang J., Tyan S.H., Chen S., et al. The ALS-FTD-linked gene product, C9orf72, regulates neuronal morphogenesis via autophagy. Autophagy. 2019;15:827–842. doi: 10.1080/15548627.2019.1569441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Shi X., Chang C., Yokom A.L., Jensen L.E., Hurley J.H. The autophagy adaptor NDP52 and the FIP200 coiled-coil allosterically activate ULK1 complex membrane recruitment. Elife. 2020;9 doi: 10.7554/eLife.59099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Jung J., Nayak A., Schaeffer V., Starzetz T., Kirsch A.K., Müller S., et al. Multiplex image-based autophagy RNAi screening identifies SMCR8 as ULK1 kinase activity and gene expression regulator. Elife. 2017;6 doi: 10.7554/eLife.23063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Ding X., Jiang X., Tian R., Zhao P., Li L., Wang X., et al. RAB2 regulates the formation of autophagosome and autolysosome in mammalian cells. Autophagy. 2019;15:1774–1786. doi: 10.1080/15548627.2019.1596478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Tang Y., Tao Y., Wang L., Yang L., Jing Y., Jiang X., et al. NPM1 mutant maintains ULK1 protein stability via TRAF6-dependent ubiquitination to promote autophagic cell survival in leukemia. FASEB J. 2021;35 doi: 10.1096/fj.201903183RRR. [DOI] [PubMed] [Google Scholar]
  • 191.Chauhan S., Mandell M.A., Deretic V. IRGM governs the core autophagy machinery to conduct antimicrobial defense. Mol Cell. 2015;58:507–521. doi: 10.1016/j.molcel.2015.03.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.McKnight N.C., Jefferies H.B., Alemu E.A., Saunders R.E., Howell M., Johansen T., et al. Genome-wide siRNA screen reveals amino acid starvation-induced autophagy requires SCOC and WAC. EMBO J. 2012;31:1931–1946. doi: 10.1038/emboj.2012.36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Liu N., Zhao H., Zhao Y.G., Hu J., Zhang H. Atlastin 2/3 regulate ER targeting of the ULK1 complex to initiate autophagy. J Cell Biol. 2021;220 doi: 10.1083/jcb.202012091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Saha B., Chisholm D., Kell A.M., Mandell M.A. A non-canonical role for the autophagy machinery in anti-retroviral signaling mediated by TRIM5α. PLoS Pathog. 2020;16 doi: 10.1371/journal.ppat.1009017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Mandell M.A., Jain A., Arko-Mensah J., Chauhan S., Kimura T., Dinkins C., et al. TRIM proteins regulate autophagy and can target autophagic substrates by direct recognition. Dev Cell. 2014;30:394–409. doi: 10.1016/j.devcel.2014.06.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.New M., Van Acker T., Sakamaki J.I., Jiang M., Saunders R.E., Long J., et al. MDH1 and MPP7 regulate autophagy in pancreatic ductal adenocarcinoma. Cancer Res. 2019;79:1884–1898. doi: 10.1158/0008-5472.CAN-18-2553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Wu J., Gao F., Xu T., Li J., Hu Z., Wang C., et al. CLDN1 induces autophagy to promote proliferation and metastasis of esophageal squamous carcinoma through AMPK/STAT1/ULK1 signaling. J Cell Physiol. 2020;235:2245–2259. doi: 10.1002/jcp.29133. [DOI] [PubMed] [Google Scholar]
  • 198.Dewi F.R.P., Jiapaer S., Kobayashi A., Hazawa M., Ikliptikawati D.K., Hartono, et al. Nucleoporin TPR (translocated promoter region, nuclear basket protein) upregulation alters MTOR–HSF1 trails and suppresses autophagy induction in ependymoma. Autophagy. 2021;17:1001–1012. doi: 10.1080/15548627.2020.1741318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Wu J., Zhang D., Li J., Deng X., Liang G., Long Y., et al. MACC1 induces autophagy to regulate proliferation, apoptosis, migration and invasion of squamous cell carcinoma. Oncol Rep. 2017;38:2369–2377. doi: 10.3892/or.2017.5889. [DOI] [PubMed] [Google Scholar]
  • 200.Han H., Yang C., Ma J., Zhang S., Zheng S., Ling R., et al. N7-Methylguanosine tRNA modification promotes esophageal squamous cell carcinoma tumorigenesis via the RPTOR/ULK1/autophagy axis. Nat Commun. 2022;13:1478. doi: 10.1038/s41467-022-29125-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Zhang S., Zhang J., An Y., Zeng X., Qin Z., Zhao Y., et al. Multi-omics approaches identify SF3B3 and SIRT3 as candidate autophagic regulators and druggable targets in invasive breast carcinoma. Acta Pharm Sin B. 2021;11:1227–1245. doi: 10.1016/j.apsb.2020.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Deng S., Shanmugam M.K., Kumar A.P., Yap C.T., Sethi G., Bishayee A. Targeting autophagy using natural compounds for cancer prevention and therapy. Cancer. 2019;125:1228–1246. doi: 10.1002/cncr.31978. [DOI] [PubMed] [Google Scholar]
  • 203.Li Y., Wang C., Xu T., Pan P., Yu Q., Xu L., et al. Discovery of a small molecule inhibitor of cullin neddylation that triggers ER stress to induce autophagy. Acta Pharm Sin B. 2021;11:3567–3584. doi: 10.1016/j.apsb.2021.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Tang J., Deng R., Luo R.Z., Shen G.P., Cai M.Y., Du Z.M., et al. Low expression of ULK1 is associated with operable breast cancer progression and is an adverse prognostic marker of survival for patients. Breast Cancer Res Treat. 2012;134:549–560. doi: 10.1007/s10549-012-2080-y. [DOI] [PubMed] [Google Scholar]
  • 205.Li W., Tanikawa T., Kryczek I., Xia H., Li G., Wu K., et al. Aerobic glycolysis controls myeloid-derived suppressor cells and tumor immunity via a specific CEBPB isoform in triple-negative breast cancer. Cell Metab. 2018;28:87–103.e6. doi: 10.1016/j.cmet.2018.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Xie C.M., Liu X.Y., Sham K.W., Lai J.M., Cheng C.H. Silencing of EEF2K (eukaryotic elongation factor-2 kinase) reveals AMPK–ULK1-dependent autophagy in colon cancer cells. Autophagy. 2014;10:1495–1508. doi: 10.4161/auto.29164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Qi W., Zhou X., Wang J., Zhang K., Zhou Y., Chen S., et al. Cordyceps sinensis polysaccharide inhibits colon cancer cells growth by inducing apoptosis and autophagy flux blockage via mTOR signaling. Carbohydr Polym. 2020;237 doi: 10.1016/j.carbpol.2020.116113. [DOI] [PubMed] [Google Scholar]
  • 208.Zhang P., Lai Z.L., Chen H.F., Zhang M., Wang A., Jia T., et al. Curcumin synergizes with 5-fluorouracil by impairing AMPK/ULK1-dependent autophagy, AKT activity and enhancing apoptosis in colon cancer cells with tumor growth inhibition in xenograft mice. J Exp Clin Cancer Res. 2017;36:190. doi: 10.1186/s13046-017-0661-7. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 209.Lee D.E., Yoo J.E., Kim J., Kim S., Kim S., Lee H., et al. NEDD4L downregulates autophagy and cell growth by modulating ULK1 and a glutamine transporter. Cell Death Dis. 2020;11:38. doi: 10.1038/s41419-020-2242-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Kinsey C.G., Camolotto S.A., Boespflug A.M., Guillen K.P., Foth M., Truong A., et al. Protective autophagy elicited by RAF→MEK→ERK inhibition suggests a treatment strategy for RAS-driven cancers. Nat Med. 2019;25:620–627. doi: 10.1038/s41591-019-0367-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Wu D.H., Wang T.T., Ruan D.Y., Li X., Chen Z.H., Wen J.Y., et al. Combination of ULK1 and LC3B improve prognosis assessment of hepatocellular carcinoma. Biomed Pharmacother. 2018;97:195–202. doi: 10.1016/j.biopha.2017.10.025. [DOI] [PubMed] [Google Scholar]
  • 212.Shin J.H., Park C.W., Yoon G., Hong S.M., Choi K.Y. NNMT depletion contributes to liver cancer cell survival by enhancing autophagy under nutrient starvation. Oncogenesis. 2018;7:58. doi: 10.1038/s41389-018-0064-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Marin J.J.G., Lozano E., Perez M.J. Lack of mitochondrial DNA impairs chemical hypoxia-induced autophagy in liver tumor cells through ROS–AMPK–ULK1 signaling dysregulation independently of HIF-1α. Free Radic Biol Med. 2016;101:71–84. doi: 10.1016/j.freeradbiomed.2016.09.025. [DOI] [PubMed] [Google Scholar]
  • 214.Xue S.T., Li K., Gao Y., Zhao L.Y., Gao Y., Yi H., et al. The role of the key autophagy kinase ULK1 in hepatocellular carcinoma and its validation as a treatment target. Autophagy. 2020;16:1823–1837. doi: 10.1080/15548627.2019.1709762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Blessing A.M., Rajapakshe K., Reddy Bollu L., Shi Y., White M.A., Pham A.H., et al. Transcriptional regulation of core autophagy and lysosomal genes by the androgen receptor promotes prostate cancer progression. Autophagy. 2017;13:506–521. doi: 10.1080/15548627.2016.1268300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Hwang D.Y., Eom J.I., Jang J.E., Jeung H.K., Chung H., Kim J.S., et al. ULK1 inhibition as a targeted therapeutic strategy for FLT3-ITD-mutated acute myeloid leukemia. J Exp Clin Cancer Res. 2020;39:85. doi: 10.1186/s13046-020-01580-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Yang W., Li Y., Liu S., Sun W., Huang H., Zhang Q., et al. Inhibition of ULK1 promotes the death of leukemia cell in an autophagy irrelevant manner and exerts the antileukemia effect. Clin Transl Med. 2021;11:e282. doi: 10.1002/ctm2.282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Ianniciello A., Zarou M.M., Rattigan K.M., Scott M., Dawson A., Dunn K., et al. ULK1 inhibition promotes oxidative stress-induced differentiation and sensitizes leukemic stem cells to targeted therapy. Sci Transl Med. 2021;13 doi: 10.1126/scitranslmed.abd5016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Obba S., Hizir Z., Boyer L., Selimoglu-Buet D., Pfeifer A., Michel G., et al. The PRKAA1/AMPKα1 pathway triggers autophagy during CSF1-induced human monocyte differentiation and is a potential target in CMML. Autophagy. 2015;11:1114–1129. doi: 10.1080/15548627.2015.1034406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Gammoh N., Fraser J., Puente C., Syred H.M., Kang H., Ozawa T., et al. Suppression of autophagy impedes glioblastoma development and induces senescence. Autophagy. 2016;12:1431–1439. doi: 10.1080/15548627.2016.1190053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Chen M.B., Ji X.Z., Liu Y.Y., Zeng P., Xu X.Y., Ma R., et al. Ulk1 over-expression in human gastric cancer is correlated with patients' T classification and cancer relapse. Oncotarget. 2017;8:33704–33712. doi: 10.18632/oncotarget.16734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Lu J., Zhu L., Zheng L.P., Cui Q., Zhu H.H., Zhao H., et al. Overexpression of ULK1 represents a potential diagnostic marker for clear cell renal carcinoma and the antitumor effects of SBI-0206965. EBioMedicine. 2018;34:85–93. doi: 10.1016/j.ebiom.2018.07.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Liu L., Yan L., Liao N., Wu W.Q., Shi J.L. A review of ULK1-mediated autophagy in drug resistance of cancer. Cancers. 2020;12:352. doi: 10.3390/cancers12020352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Jang J.E., Eom J.I., Jeung H.K., Cheong J.W., Lee J.Y., Kim J.S., et al. AMPK–ULK1-mediated autophagy confers resistance to BET inhibitor JQ1 in acute myeloid leukemia stem cells. Clin Cancer Res. 2017;23:2781–2794. doi: 10.1158/1078-0432.CCR-16-1903. [DOI] [PubMed] [Google Scholar]
  • 225.Ianniciello A., Helgason G.V. Targeting ULK1 in cancer stem cells: insight from chronic myeloid leukemia. Autophagy. 2022;18:1734–1736. doi: 10.1080/15548627.2022.2041152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Huang J., Ni J., Liu K., Yu Y., Xie M., Kang R., et al. HMGB1 promotes drug resistance in osteosarcoma. Cancer Res. 2012;72:230–238. doi: 10.1158/0008-5472.CAN-11-2001. [DOI] [PubMed] [Google Scholar]
  • 227.Chen W., Hu Y., Ju D. Gene therapy for neurodegenerative disorders: advances, insights and prospects. Acta Pharm Sin B. 2020;10:1347–1359. doi: 10.1016/j.apsb.2020.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Fan L., Qiu X.X., Zhu Z.Y., Lv J.L., Lu J., Mao F., et al. Nitazoxanide, an anti-parasitic drug, efficiently ameliorates learning and memory impairments in AD model mice. Acta Pharmacol Sin. 2019;40:1279–1291. doi: 10.1038/s41401-019-0220-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Zhang W., Xu C., Sun J., Shen H.M., Wang J., Yang C. Impairment of the autophagy-lysosomal pathway in Alzheimer's diseases: pathogenic mechanisms and therapeutic potential. Acta Pharm Sin B. 2022;12:1019–1040. doi: 10.1016/j.apsb.2022.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Miki Y., Shimoyama S., Kon T., Ueno T., Hayakari R., Tanji K., et al. Alteration of autophagy-related proteins in peripheral blood mononuclear cells of patients with Parkinson's disease. Neurobiol Aging. 2018;63:33–43. doi: 10.1016/j.neurobiolaging.2017.11.006. [DOI] [PubMed] [Google Scholar]
  • 231.Zhang K., Zhu S., Li J., Jiang T., Feng L., Pei J., et al. Targeting autophagy using small-molecule compounds to improve potential therapy of Parkinson's disease. Acta Pharm Sin B. 2021;11:3015–3034. doi: 10.1016/j.apsb.2021.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Ouyang L., Zhang L., Zhang S., Yao D., Zhao Y., Wang G., et al. Small-molecule activator of UNC-51-like kinase 1 (ULK1) that induces cytoprotective autophagy for Parkinson's disease treatment. J Med Chem. 2018;61:2776–2792. doi: 10.1021/acs.jmedchem.7b01575. [DOI] [PubMed] [Google Scholar]
  • 233.Franco-Iborra S., Plaza-Zabala A., Montpeyo M., Sebastian D., Vila M., Martinez-Vicente M. Mutant HTT (huntingtin) impairs mitophagy in a cellular model of Huntington disease. Autophagy. 2021;17:672–689. doi: 10.1080/15548627.2020.1728096. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Webster C.P., Smith E.F., Bauer C.S., Moller A., Hautbergue G.M., Ferraiuolo L., et al. The C9orf72 protein interacts with Rab1a and the ULK1 complex to regulate initiation of autophagy. EMBO J. 2016;35:1656–1676. doi: 10.15252/embj.201694401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Deng Z., Lim J., Wang Q., Purtell K., Wu S., Palomo G.M., et al. ALS-FTLD-linked mutations of SQSTM1/p62 disrupt selective autophagy and NFE2L2/NRF2 anti-oxidative stress pathway. Autophagy. 2020;16:917–931. doi: 10.1080/15548627.2019.1644076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Lupo F., Tibaldi E., Matte A., Sharma A.K., Brunati A.M., Alper S.L., et al. A new molecular link between defective autophagy and erythroid abnormalities in chorea-acanthocytosis. Blood. 2016;128:2976–2987. doi: 10.1182/blood-2016-07-727321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Ribas V.T., Vahsen B.F., Tatenhorst L., Estrada V., Dambeck V., Almeida R.A., et al. AAV-mediated inhibition of ULK1 promotes axonal regeneration in the central nervous system in vitro and in vivo. Cell Death Dis. 2021;12:213. doi: 10.1038/s41419-021-03503-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Wei H.L., Ma S.Q., Li C.X. Deficiency of unc-51 like kinase 1 (Ulk1) protects against mice traumatic brain injury (TBI) by suppression of p38 and JNK pathway. Biochem Biophys Res Commun. 2018;503:467–473. doi: 10.1016/j.bbrc.2018.04.154. [DOI] [PubMed] [Google Scholar]
  • 239.She H., He Y., Zhao Y., Mao Z. Autophagy in inflammation: the p38α MAPK–ULK1 axis. Macrophage (Houst) 2018;5:e1629. [PMC free article] [PubMed] [Google Scholar]
  • 240.Li Z., Zhao K., Lv X., Lan Y., Hu S., Shi J., et al. Ulk1 governs nerve growth factor/TrkA signaling by mediating Rab5 GTPase activation in porcine hemagglutinating encephalomyelitis virus-induced neurodegenerative disorders. J Virol. 2018;92 doi: 10.1128/JVI.00325-18. e00325-18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Al Eissa M.M., Fiorentino A., Sharp S.I., O'Brien N.L., Wolfe K., Giaroli G., et al. Exome sequence analysis and follow up genotyping implicates rare ULK1 variants to be involved in susceptibility to schizophrenia. Ann Hum Genet. 2018;82:88–92. doi: 10.1111/ahg.12226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Ferreira da Silva T., Granadeiro L.S., Bessa-Neto D., Luz L.L., Safronov B.V., Brites P. Plasmalogens regulate the AKT–ULK1 signaling pathway to control the position of the axon initial segment. Prog Neurobiol. 2021;205 doi: 10.1016/j.pneurobio.2021.102123. [DOI] [PubMed] [Google Scholar]
  • 243.Wang X., Gao Y., Tan J., Devadas K., Ragupathy V., Takeda K., et al. HIV-1 and HIV-2 infections induce autophagy in Jurkat and CD4+ T cells. Cell Signal. 2012;24:1414–1419. doi: 10.1016/j.cellsig.2012.02.016. [DOI] [PubMed] [Google Scholar]
  • 244.Nozawa T., Sano S., Minowa-Nozawa A., Toh H., Nakajima S., Murase K., et al. TBC1D9 regulates TBK1 activation through Ca2+ signaling in selective autophagy. Nat Commun. 2020;11:770. doi: 10.1038/s41467-020-14533-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Rubio R.M., Mohr I. Inhibition of ULK1 and Beclin1 by an α-herpesvirus Akt-like Ser/Thr kinase limits autophagy to stimulate virus replication. Proc Natl Acad Sci U S A. 2019;116:26941–26950. doi: 10.1073/pnas.1915139116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Owen K.A., Meyer C.B., Bouton A.H., Casanova J.E. Activation of focal adhesion kinase by Salmonella suppresses autophagy via an Akt/mTOR signaling pathway and promotes bacterial survival in macrophages. PLoS Pathog. 2014;10 doi: 10.1371/journal.ppat.1004159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Feng Z.Z., Jiang A.J., Mao A.W., Feng Y., Wang W., Li J., et al. The Salmonella effectors SseF and SseG inhibit Rab1A-mediated autophagy to facilitate intracellular bacterial survival and replication. J Biol Chem. 2018;293:9662–9673. doi: 10.1074/jbc.M117.811737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Starr T., Child R., Wehrly T.D., Hansen B., Hwang S., López-Otin C., et al. Selective subversion of autophagy complexes facilitates completion of the Brucella intracellular cycle. Cell Host Microbe. 2012;11:33–45. doi: 10.1016/j.chom.2011.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Hansen M.D., Johnsen I.B., Stiberg K.A., Sherstova T., Wakita T., Richard G.M., et al. Hepatitis C virus triggers Golgi fragmentation and autophagy through the immunity-related GTPase M. Proc Natl Acad Sci U S A. 2017;114 doi: 10.1073/pnas.1616683114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Pandey A., Ding S.L., Qin Q.M., Gupta R., Gomez G., Lin F., et al. Global reprogramming of host kinase signaling in response to fungal infection. Cell Host Microbe. 2017;21:637–649.e6. doi: 10.1016/j.chom.2017.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Wang X., Lin Y., Kemper T., Chen J., Yuan Z., Liu S., et al. AMPK and Akt/mTOR signalling pathways participate in glucose-mediated regulation of hepatitis B virus replication and cellular autophagy. Cell Microbiol. 2020;22 doi: 10.1111/cmi.13131. [DOI] [PubMed] [Google Scholar]
  • 252.Zhao C., Chen J., Cheng L., Xu K., Yang Y., Su X. Deficiency of HIF-1α enhances influenza A virus replication by promoting autophagy in alveolar type II epithelial cells. Emerg Microbes Infect. 2020;9:691–706. doi: 10.1080/22221751.2020.1742585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Saleiro D., Kosciuczuk E.M., Platanias L.C. Beyond autophagy: new roles for ULK1 in immune signaling and interferon responses. Cytokine Growth Factor Rev. 2016;29:17–22. doi: 10.1016/j.cytogfr.2016.03.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Saleiro D., Mehrotra S., Kroczynska B., Beauchamp E.M., Lisowski P., Majchrzak-Kita B., et al. Central role of ULK1 in type I interferon signaling. Cell Rep. 2015;11:605–617. doi: 10.1016/j.celrep.2015.03.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Saleiro D., Blyth G.T., Kosciuczuk E.M., Ozark P.A., Majchrzak-Kita B., Arslan A.D., et al. IFN-γ-inducible antiviral responses require ULK1-mediated activation of MLK3 and ERK5. Sci Signal. 2018;11 doi: 10.1126/scisignal.aap9921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Saul V.V., Seibert M., Krüger M., Jeratsch S., Kracht M., Schmitz M.L. ULK1/2 restricts the formation of inducible SINT-speckles, membraneless organelles controlling the threshold of TBK1 activation. iScience. 2019;19:527–544. doi: 10.1016/j.isci.2019.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Shi J.H., Ling C., Wang T.T., Zhang L.N., Liu W.W., Qin Y., et al. TRK-fused gene (TFG) regulates ULK1 stability via TRAF3-mediated ubiquitination and protects macrophages from LPS-induced pyroptosis. Cell Death Dis. 2022;13:93. doi: 10.1038/s41419-022-04539-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Wang J., Zhou H. Mitochondrial quality control mechanisms as molecular targets in cardiac ischemia-reperfusion injury. Acta Pharm Sin B. 2020;10:1866–1879. doi: 10.1016/j.apsb.2020.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Oduro P.K., Zheng X., Wei J., Yang Y., Wang Y., Zhang H., et al. The cGAS–STING signaling in cardiovascular and metabolic diseases: future novel target option for pharmacotherapy. Acta Pharm Sin B. 2022;12:50–75. doi: 10.1016/j.apsb.2021.05.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Takagi H., Matsui Y., Hirotani S., Sakoda H., Asano T., Sadoshima J. AMPK mediates autophagy during myocardial ischemia in vivo. Autophagy. 2007;3:405–407. doi: 10.4161/auto.4281. [DOI] [PubMed] [Google Scholar]
  • 261.Sung H.K., Chan Y.K., Han M., Jahng J.W.S., Song E., Danielson E., et al. Lipocalin-2 (NGAL) attenuates autophagy to exacerbate cardiac apoptosis induced by myocardial ischemia. J Cell Physiol. 2017;232:2125–2134. doi: 10.1002/jcp.25672. [DOI] [PubMed] [Google Scholar]
  • 262.Zheng D., Liu Z., Zhou Y., Hou N., Yan W., Qin Y., et al. Urolithin B, a gut microbiota metabolite, protects against myocardial ischemia/reperfusion injury via p62/Keap1/Nrf2 signaling pathway. Pharmacol Res. 2020;153 doi: 10.1016/j.phrs.2020.104655. [DOI] [PubMed] [Google Scholar]
  • 263.An M., Ryu D.R., Won Park J., Ha Choi J., Park E.M., Eun Lee K., et al. ULK1 prevents cardiac dysfunction in obesity through autophagy-meditated regulation of lipid metabolism. Cardiovasc Res. 2017;113:1137–1147. doi: 10.1093/cvr/cvx064. [DOI] [PubMed] [Google Scholar]
  • 264.Chen C., Jiang L., Zhang M., Pan X., Peng C., Huang W., et al. Isodunnianol alleviates doxorubicin-induced myocardial injury by activating protective autophagy. Food Funct. 2019;10:2651–2657. doi: 10.1039/c9fo00063a. [DOI] [PubMed] [Google Scholar]
  • 265.Wang L., Yuan D., Zheng J., Wu X., Wang J., Liu X., et al. Chikusetsu saponin IVa attenuates isoprenaline-induced myocardial fibrosis in mice through activation autophagy mediated by AMPK/mTOR/ULK1 signaling. Phytomedicine. 2019;58 doi: 10.1016/j.phymed.2018.11.024. [DOI] [PubMed] [Google Scholar]
  • 266.Yu W., Zha W., Ren J. Exendin-4 and liraglutide attenuate glucose toxicity-induced cardiac injury through mTOR/ULK1-dependent autophagy. Oxid Med Cell Longev. 2018;2018 doi: 10.1155/2018/5396806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Baranyai T., Nagy C.T., Koncsos G., Onódi Z., Károlyi-Szabó M., Makkos A., et al. Acute hyperglycemia abolishes cardioprotection by remote ischemic perconditioning. Cardiovasc Diabetol. 2015;14:151. doi: 10.1186/s12933-015-0313-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Guo R., Ren J. Deficiency in AMPK attenuates ethanol-induced cardiac contractile dysfunction through inhibition of autophagosome formation. Cardiovasc Res. 2012;94:480–491. doi: 10.1093/cvr/cvs127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Jiang S., Guo R., Zhang Y., Zou Y., Ren J. Heavy metal scavenger metallothionein mitigates deep hypothermia-induced myocardial contractile anomalies: role of autophagy. Am J Physiol Endocrinol Metab. 2013;304:E74–E86. doi: 10.1152/ajpendo.00176.2012. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 270.Fiuza-Luces C., Delmiro A., Soares-Miranda L., González-Murillo Á., Martínez-Palacios J., Ramírez M., et al. Exercise training can induce cardiac autophagy at end-stage chronic conditions: insights from a graft-versus-host-disease mouse model. Brain Behav Immun. 2014;39:56–60. doi: 10.1016/j.bbi.2013.11.007. [DOI] [PubMed] [Google Scholar]
  • 271.Steiner J.L., Gordon B.S., Lang C.H. Moderate alcohol consumption does not impair overload-induced muscle hypertrophy and protein synthesis. Physiol Rep. 2015;3 doi: 10.14814/phy2.12333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Nah J., Shirakabe A., Mukai R., Zhai P., Sung E.A., Ivessa A., et al. Ulk1-dependent alternative mitophagy plays a protective role during pressure overload in the heart. Cardiovasc Res. 2022 doi: 10.1093/cvr/cvac003. Available from: [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Tong M., Saito T., Zhai P., Oka S.I., Mizushima W., Nakamura M., et al. Alternative mitophagy protects the heart against obesity-associated cardiomyopathy. Circ Res. 2021;129:1105–1121. doi: 10.1161/CIRCRESAHA.121.319377. [DOI] [PubMed] [Google Scholar]
  • 274.Harris M.P., Zhang Q.J., Cochran C.T., Ponce J., Alexander S., Kronemberger A., et al. Perinatal versus adult loss of ULK1 and ULK2 distinctly influences cardiac autophagy and function. Autophagy. 2022;18:2161–2177. doi: 10.1080/15548627.2021.2022289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Zhang A., Wang M., Zhuo P. Unc-51 like autophagy activating kinase 1 accelerates angiotensin II-induced cardiac hypertrophy through promoting oxidative stress regulated by Nrf-2/HO-1 pathway. Biochem Biophys Res Commun. 2019;509:32–39. doi: 10.1016/j.bbrc.2018.11.190. [DOI] [PubMed] [Google Scholar]
  • 276.Gao Y., Zhang W., Zeng L.Q., Bai H., Li J., Zhou J., et al. Exercise and dietary intervention ameliorate high-fat diet-induced NAFLD and liver aging by inducing lipophagy. Redox Biol. 2020;36 doi: 10.1016/j.redox.2020.101635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Shimizu S., Yoshida T., Tsujioka M., Arakawa S. Autophagic cell death and cancer. Int J Mol Sci. 2014;15:3145–3153. doi: 10.3390/ijms15023145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Yang L., Wu Y., Lin S., Dai B., Chen H., Tao X., et al. sPLA2-IB and PLA2R mediate insufficient autophagy and contribute to podocyte injury in idiopathic membranous nephropathy by activation of the p38MAPK/mTOR/ULK1(ser757) signaling pathway. FASEB J. 2021;35 doi: 10.1096/fj.202001143R. [DOI] [PubMed] [Google Scholar]
  • 279.Zhang Y., Zhang S., Wang Y., Yang Z., Chen Z., Wen N., et al. ULK1 suppresses osteoclast differentiation and bone resorption via inhibiting Syk-JNK through DOK3. Oxid Med Cell Longev. 2021;2021 doi: 10.1155/2021/2896674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Zhao H., Sun Z., Ma Y., Song R., Yuan Y., Bian J., et al. Antiosteoclastic bone resorption activity of osteoprotegerin via enhanced AKT/mTOR/ULK1-mediated autophagic pathway. J Cell Physiol. 2020;235:3002–3012. doi: 10.1002/jcp.29205. [DOI] [PubMed] [Google Scholar]
  • 281.Chen X., Li L., Xu S., Bu W., Chen K., Li M., et al. Ultraviolet B radiation down-regulates ULK1 and ATG7 expression and impairs the autophagy response in human keratinocytes. J Photochem Photobiol B. 2018;178:152–164. doi: 10.1016/j.jphotobiol.2017.08.043. [DOI] [PubMed] [Google Scholar]
  • 282.Shibata S., Ishizawa K., Wang Q., Xu N., Fujita T., Uchida S., et al. ULK1 phosphorylates and regulates mineralocorticoid receptor. Cell Rep. 2018;24:569–576. doi: 10.1016/j.celrep.2018.06.072. [DOI] [PubMed] [Google Scholar]
  • 283.Su Y., Lu J., Chen X., Liang C., Luo P., Qin C., et al. Rapamycin alleviates hormone imbalance-induced chronic nonbacterial inflammation in rat prostate through activating autophagy via the mTOR/ULK1/ATG13 signaling pathway. Inflammation. 2018;41:1384–1395. doi: 10.1007/s10753-018-0786-7. [DOI] [PubMed] [Google Scholar]
  • 284.Sun Y., Li T.Y., Song L., Zhang C., Li J., Lin Z.Z., et al. Liver-specific deficiency of unc-51 like kinase 1 and 2 protects mice from acetaminophen-induced liver injury. Hepatology. 2018;67:2397–2413. doi: 10.1002/hep.29759. [DOI] [PubMed] [Google Scholar]
  • 285.Sinha R.A., Singh B.K., Zhou J., Xie S., Farah B.L., Lesmana R., et al. Loss of ULK1 increases RPS6KB1-NCOR1 repression of NR1H/LXR-mediated Scd1 transcription and augments lipotoxicity in hepatic cells. Autophagy. 2017;13:169–186. doi: 10.1080/15548627.2016.1235123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Wu W., Stork B. Regulating RIPK1: another way in which ULK1 contributes to survival. Autophagy. 2020;16:1544–1546. doi: 10.1080/15548627.2020.1783110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Lazarus M.B., Shokat K.M. Discovery and structure of a new inhibitor scaffold of the autophagy initiating kinase ULK1. Bioorg Med Chem. 2015;23:5483–5488. doi: 10.1016/j.bmc.2015.07.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Petherick K.J., Conway O.J., Mpamhanga C., Osborne S.A., Kamal A., Saxty B., et al. Pharmacological inhibition of ULK1 kinase blocks mammalian target of rapamycin (mTOR)-dependent autophagy. J Biol Chem. 2015;290:11376–11383. doi: 10.1074/jbc.C114.627778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Chen Y., Xie X., Wang C., Hu Y., Zhang H., Zhang L., et al. Dual targeting of NUAK1 and ULK1 using the multitargeted inhibitor MRT68921 exerts potent antitumor activities. Cell Death Dis. 2020;11:712. doi: 10.1038/s41419-020-02885-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Carta D., Bortolozzi R., Hamel E., Basso G., Moro S., Viola G., et al. Novel 3-substituted 7-phenylpyrrolo[3,2-f]quinolin-9(6H)-ones as single entities with multitarget antiproliferative activity. J Med Chem. 2015;58:7991–8010. doi: 10.1021/acs.jmedchem.5b00805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Wood S.D., Grant W., Adrados I., Choi J.Y., Alburger J.M., Duckett D.R., et al. In silico HTS and structure based optimization of indazole-derived ULK1 inhibitors. ACS Med Chem Lett. 2017;8:1258–1263. doi: 10.1021/acsmedchemlett.7b00344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Knudsen J.R., Madsen A.B., Persson K.W., Henríquez-Olguín C., Li Z., Jensen T.E. The ULK1/2 and AMPK inhibitor SBI-0206965 blocks AICAR and insulin-stimulated glucose transport. Int J Mol Sci. 2020;21:2344. doi: 10.3390/ijms21072344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Ren H., Bakas N.A., Vamos M., Chaikuad A., Limpert A.S., Wimer C.D., et al. Design, synthesis, and characterization of an orally active dual-specific ULK1/2 autophagy inhibitor that synergizes with the PARP inhibitor olaparib for the treatment of triple-negative breast cancer. J Med Chem. 2020;63:14609–14625. doi: 10.1021/acs.jmedchem.0c00873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Sun D., Yang Z., Zhen Y., Yang Y., Chen Y., Yuan Y., et al. Discovery of 5-bromo-4-phenoxy-N-phenylpyrimidin-2-amine derivatives as novel ULK1 inhibitors that block autophagy and induce apoptosis in non-small cell lung cancer. Eur J Med Chem. 2020;208 doi: 10.1016/j.ejmech.2020.112782. [DOI] [PubMed] [Google Scholar]
  • 295.Xiang H., Zhang J., Lin C., Zhang L., Liu B., Ouyang L. Targeting autophagy-related protein kinases for potential therapeutic purpose. Acta Pharm Sin B. 2020;10:569–581. doi: 10.1016/j.apsb.2019.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Zhang Y., Miao J.M. Ginkgolide K promotes astrocyte proliferation and migration after oxygen-glucose deprivation via inducing protective autophagy through the AMPK/mTOR/ULK1 signaling pathway. Eur J Pharmacol. 2018;832:96–103. doi: 10.1016/j.ejphar.2018.05.029. [DOI] [PubMed] [Google Scholar]
  • 297.Chen W.R., Yang J.Q., Liu F., Shen X.Q., Zhou Y.J. Melatonin attenuates vascular calcification by activating autophagy via an AMPK/mTOR/ULK1 signaling pathway. Exp Cell Res. 2020;389 doi: 10.1016/j.yexcr.2020.111883. [DOI] [PubMed] [Google Scholar]
  • 298.Gui D., Cui Z., Zhang L., Yu C., Yao D., Xu M., et al. Salidroside attenuates hypoxia-induced pulmonary arterial smooth muscle cell proliferation and apoptosis resistance by upregulating autophagy through the AMPK–mTOR–ULK1 pathway. BMC Pulm Med. 2017;17:191. doi: 10.1186/s12890-017-0477-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Wu C., Jing M., Yang L., Jin L., Ding Y., Lu J., et al. Alisol A 24-acetate ameliorates nonalcoholic steatohepatitis by inhibiting oxidative stress and stimulating autophagy through the AMPK/mTOR pathway. Chem Biol Interact. 2018;291:111–119. doi: 10.1016/j.cbi.2018.06.005. [DOI] [PubMed] [Google Scholar]
  • 300.Jin Y., Liu S., Ma Q., Xiao D., Chen L. Berberine enhances the AMPK activation and autophagy and mitigates high glucose-induced apoptosis of mouse podocytes. Eur J Pharmacol. 2017;794:106–114. doi: 10.1016/j.ejphar.2016.11.037. [DOI] [PubMed] [Google Scholar]
  • 301.Wang F., Cao M., Fan M., Wu H., Huang W., Zhang Y., et al. AMPK–mTOR–ULK1 axis activation-dependent autophagy promotes hydroxycamptothecin-induced apoptosis in human bladder cancer cells. J Cell Physiol. 2020;235:4302–4315. doi: 10.1002/jcp.29307. [DOI] [PubMed] [Google Scholar]
  • 302.Ouyang L., Zhang L., Liu J., Fu L., Yao D., Zhao Y., et al. Discovery of a small-molecule bromodomain-containing protein 4 (BRD4) inhibitor that induces AMP-Activated protein kinase-modulated autophagy-associated cell death in breast cancer. J Med Chem. 2017;60:9990–10012. doi: 10.1021/acs.jmedchem.7b00275. [DOI] [PubMed] [Google Scholar]
  • 303.Cai B., Gong L., Zhu Y., Kong L., Ju X., Li X., et al. Identification of gossypol acetate as an autophagy modulator with potent anti-tumor effect against cancer cells. J Agric Food Chem. 2022;70:2589–2599. doi: 10.1021/acs.jafc.1c06399. [DOI] [PubMed] [Google Scholar]
  • 304.Yang S., Long L.H., Li D., Zhang J.K., Jin S., Wang F., et al. β-Guanidinopropionic acid extends the lifespan of Drosophila melanogaster via an AMP-activated protein kinase-dependent increase in autophagy. Aging Cell. 2015;14:1024–1033. doi: 10.1111/acel.12371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Kim S.H., Park S., Yu H.S., Ko K.H., Park H.G., Kim Y.S. The antipsychotic agent clozapine induces autophagy via the AMPK–ULK1–Beclin1 signaling pathway in the rat frontal cortex. Prog Neuro-Psychopharmacol Biol Psychiatry. 2018;81:96–104. doi: 10.1016/j.pnpbp.2017.10.012. [DOI] [PubMed] [Google Scholar]
  • 306.Kim S.W., Rho C.R., Kim J., Xie Y., Prince R.C., Mustafa K., et al. Eicosapentaenoic acid (EPA) activates PPARγ signaling leading to cell cycle exit, lipid accumulation, and autophagy in human meibomian gland epithelial cells (hMGEC) Ocul Surf. 2020;18:427–437. doi: 10.1016/j.jtos.2020.04.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Yu J., Li X., Matei N., McBride D., Tang J., Yan M., et al. Ezetimibe, a NPC1L1 inhibitor, attenuates neuronal apoptosis through AMPK dependent autophagy activation after MCAO in rats. Exp Neurol. 2018;307:12–23. doi: 10.1016/j.expneurol.2018.05.022. [DOI] [PubMed] [Google Scholar]
  • 308.Fan Y., Wang N., Rocchi A., Zhang W., Vassar R., Zhou Y., et al. Identification of natural products with neuronal and metabolic benefits through autophagy induction. Autophagy. 2017;13:41–56. doi: 10.1080/15548627.2016.1240855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Suvorova, Knyazeva A.R., Petukhov A.V., Aksenov N.D., Pospelov V.A. Resveratrol enhances pluripotency of mouse embryonic stem cells by activating AMPK/Ulk1 pathway. Cell Death Discov. 2019;5:61. doi: 10.1038/s41420-019-0137-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Sun J., Mu Y., Jiang Y., Song R., Yi J., Zhou J., et al. Inhibition of p70 S6 kinase activity by A77 1726 induces autophagy and enhances the degradation of superoxide dismutase 1 (SOD1) protein aggregates. Cell Death Dis. 2018;9:407. doi: 10.1038/s41419-018-0441-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Gao L., Chen X., Fu Z., Yin J., Wang Y., Sun W., et al. Kinsenoside alleviates alcoholic liver injury by reducing oxidative stress, inhibiting endoplasmic reticulum stress, and regulating AMPK-dependent autophagy. Front Pharmacol. 2021;12 doi: 10.3389/fphar.2021.747325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Aryal P., Kim K., Park P.H., Ham S., Cho J., Song K. Baicalein induces autophagic cell death through AMPK/ULK1 activation and downregulation of mTORC1 complex components in human cancer cells. FEBS J. 2014;281:4644–4658. doi: 10.1111/febs.12969. [DOI] [PubMed] [Google Scholar]
  • 313.Wang Y.F., Li T., Tang Z.H., Chang L.L., Zhu H., Chen X.P., et al. Baicalein triggers autophagy and inhibits the protein kinase B/mammalian target of rapamycin pathway in hepatocellular carcinoma HepG2 Cells. Phytother Res. 2015;29:674–679. doi: 10.1002/ptr.5298. [DOI] [PubMed] [Google Scholar]
  • 314.Cao C., Huang W., Zhang N., Wu F., Xu T., Pan X., et al. Narciclasine induces autophagy-dependent apoptosis in triple-negative breast cancer cells by regulating the AMPK–ULK1 axis. Cell Prolif. 2018;51 doi: 10.1111/cpr.12518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Kim D.E., Kim Y., Cho D.H., Jeong S.Y., Kim S.B., Suh N., et al. Raloxifene induces autophagy-dependent cell death in breast cancer cells via the activation of AMP-activated protein kinase. Mol Cells. 2015;38:138–144. doi: 10.14348/molcells.2015.2193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Draz H., Goldberg A.A., Titorenko V.I., Tomlinson Guns E.S., Safe S.H., Sanderson J.T. Diindolylmethane and its halogenated derivatives induce protective autophagy in human prostate cancer cells via induction of the oncogenic protein AEG-1 and activation of AMP-activated protein kinase (AMPK) Cell Signal. 2017;40:172–182. doi: 10.1016/j.cellsig.2017.09.006. [DOI] [PubMed] [Google Scholar]
  • 317.Zou Y., Wang Q., Li B., Xie B., Wang W. Temozolomide induces autophagy via ATM–AMPK–ULK1 pathways in glioma. Mol Med Rep. 2014;10:411–416. doi: 10.3892/mmr.2014.2151. [DOI] [PubMed] [Google Scholar]
  • 318.Li B., Wu G.L., Dai W., Wang G., Su H.Y., Shen X.P., et al. Aescin-induced reactive oxygen species play a pro-survival role in human cancer cells via ATM/AMPK/ULK1-mediated autophagy. Acta Pharmacol Sin. 2018;39:1874–1884. doi: 10.1038/s41401-018-0047-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Zhao Q., Wang Z., Wang Z., Wu L., Zhang W. Aspirin may inhibit angiogenesis and induce autophagy by inhibiting mTOR signaling pathway in murine hepatocarcinoma and sarcoma models. Oncol Lett. 2016;12:2804–2810. doi: 10.3892/ol.2016.5017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Huang S., Yang Z.J., Yu C., Sinicrope F.A. Inhibition of mTOR kinase by AZD8055 can antagonize chemotherapy-induced cell death through autophagy induction and down-regulation of p62/sequestosome 1. J Biol Chem. 2011;286:40002–40012. doi: 10.1074/jbc.M111.297432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Teng J.F., Qin D.L., Mei Q.B., Qiu W.Q., Pan R., Xiong R., et al. Polyphyllin VI, a saponin from Trillium tschonoskii Maxim. induces apoptotic and autophagic cell death via the ROS triggered mTOR signaling pathway in non-small cell lung cancer. Pharmacol Res. 2019;147 doi: 10.1016/j.phrs.2019.104396. [DOI] [PubMed] [Google Scholar]
  • 322.Zhang H.W., Hu J.J., Fu R.Q., Liu X., Zhang Y.H., Li J., et al. Flavonoids inhibit cell proliferation and induce apoptosis and autophagy through downregulation of PI3Kγ mediated PI3K/AKT/mTOR/p70S6K/ULK signaling pathway in human breast cancer cells. Sci Rep. 2018;8 doi: 10.1038/s41598-018-29308-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Wang J., Ren X.R., Piao H., Zhao S., Osada T., Premont R.T., et al. Niclosamide-induced Wnt signaling inhibition in colorectal cancer is mediated by autophagy. Biochem J. 2019;476:535–546. doi: 10.1042/BCJ20180385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Shou J., Wang M., Cheng X., Wang X., Zhang L., Liu Y., et al. Tizoxanide induces autophagy by inhibiting PI3K/Akt/mTOR pathway in RAW264.7 macrophage cells. Arch Pharm Res (Seoul) 2020;43:257–270. doi: 10.1007/s12272-019-01202-4. [DOI] [PubMed] [Google Scholar]
  • 325.Niu Q., Zhao W., Wang J., Li C., Yan T., Lv W., et al. LicA induces autophagy through ULK1/Atg13 and ROS pathway in human hepatocellular carcinoma cells. Int J Mol Med. 2018;41:2601–2608. doi: 10.3892/ijmm.2018.3499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Gu Y., Han J., Xue F., Xiao H., Chen L., Zhao Z., et al. 4,4′-Dimethoxychalcone protects the skin from AAPH-induced senescence and UVB-induced photoaging by activating autophagy. Food Funct. 2022;13:4114–4129. doi: 10.1039/d1fo04130d. [DOI] [PubMed] [Google Scholar]
  • 327.Tian L., Ma L., Guo E., Deng X., Ma S., Xia Q., et al. 20-Hydroxyecdysone upregulates Atg genes to induce autophagy in the Bombyx fat body. Autophagy. 2013;9:1172–1187. doi: 10.4161/auto.24731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Jiang J., Dai J., Cui H. Vitexin reverses the autophagy dysfunction to attenuate MCAO-induced cerebral ischemic stroke via mTOR/Ulk1 pathway. Biomed Pharmacother. 2018;99:583–590. doi: 10.1016/j.biopha.2018.01.067. [DOI] [PubMed] [Google Scholar]
  • 329.Drießen S., Berleth N., Friesen O., Löffler A.S., Böhler P., Hieke N., et al. Deubiquitinase inhibition by WP1130 leads to ULK1 aggregation and blockade of autophagy. Autophagy. 2015;11:1458–1470. doi: 10.1080/15548627.2015.1067359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Yao C., Ni Z., Gong C., Zhu X., Wang L., Xu Z., et al. Rocaglamide enhances NK cell-mediated killing of non-small cell lung cancer cells by inhibiting autophagy. Autophagy. 2018;14:1831–1844. doi: 10.1080/15548627.2018.1489946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Kimura T., Uesugi M., Takase K., Miyamoto N., Sawada K. Hsp90 inhibitor geldanamycin attenuates the cytotoxicity of sunitinib in cardiomyocytes via inhibition of the autophagy pathway. Toxicol Appl Pharmacol. 2017;329:282–292. doi: 10.1016/j.taap.2017.06.015. [DOI] [PubMed] [Google Scholar]
  • 332.Liu B., Tu Y., He W., Liu Y., Wu W., Fang Q., et al. Hyperoside attenuates renal aging and injury induced by d-galactose via inhibiting AMPK–ULK1 signaling-mediated autophagy. Aging (Albany N Y) 2018;10:4197–4212. doi: 10.18632/aging.101723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Lin P.H., Chiang Y.F., Shieh T.M., Chen H.Y., Shih C.K., Wang T.H., et al. Dietary compound isoliquiritigenin, an antioxidant from licorice, suppresses triple-negative breast tumor growth via apoptotic death program activation in cell and xenograft animal models. Antioxidants. 2020;9:228. doi: 10.3390/antiox9030228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Zamame Ramirez J.A., Romagnoli G.G., Falasco B.F., Gorgulho C.M., Sanzochi Fogolin C., Dos Santos D.C., et al. Blocking drug-induced autophagy with chloroquine in HCT-116 colon cancer cells enhances DC maturation and T cell responses induced by tumor cell lysate. Int Immunopharmacol. 2020;84 doi: 10.1016/j.intimp.2020.106495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Quan Y., Lei H., Wahafu W., Liu Y., Ping H., Zhang X. Inhibition of autophagy enhances the anticancer effect of enzalutamide on bladder cancer. Biomed Pharmacother. 2019;120 doi: 10.1016/j.biopha.2019.109490. [DOI] [PubMed] [Google Scholar]
  • 336.Patra T., Meyer K., Ray R.B., Ray R. A combination of AZD5363 and FH5363 induces lethal autophagy in transformed hepatocytes. Cell Death Dis. 2020;11:540. doi: 10.1038/s41419-020-02741-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Despotović A., Mirčić A., Misirlić-Denčić S., Harhaji-Trajković L., Trajković V., Zogović N., et al. Combination of ascorbic acid and menadione induces cytotoxic autophagy in human glioblastoma cells. Oxid Med Cell Longev. 2022;2022 doi: 10.1155/2022/2998132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 338.Ibañez-Cabellos J.S., Aguado C., Pérez-Cremades D., García-Giménez J.L., Bueno-Betí C., García-López E.M., et al. Extracellular histones activate autophagy and apoptosis via mTOR signaling in human endothelial cells. Biochim Biophys Acta, Mol Basis Dis. 2018;1864:3234–3246. doi: 10.1016/j.bbadis.2018.07.010. [DOI] [PubMed] [Google Scholar]
  • 339.Feng X.L., Zheng Y., Zong M.M., Hao S.S., Zhou G.F., Cao R.B., et al. The immunomodulatory functions and molecular mechanism of a new bursal heptapeptide (BP7) in immune responses and immature B cells. Vet Res. 2019;50:64. doi: 10.1186/s13567-019-0682-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 340.Liu Y., Palanivel R., Rai E., Park M., Gabor T.V., Scheid M.P., et al. Adiponectin stimulates autophagy and reduces oxidative stress to enhance insulin sensitivity during high-fat diet feeding in mice. Diabetes. 2015;64:36–48. doi: 10.2337/db14-0267. [DOI] [PubMed] [Google Scholar]
  • 341.Sehdev A., Karrison T., Zha Y., Janisch L., Turcich M., Cohen E.E.W., et al. A pharmacodynamic study of sirolimus and metformin in patients with advanced solid tumors. Cancer Chemother Pharmacol. 2018;82:309–317. doi: 10.1007/s00280-018-3619-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Zhang C., He X., Kwok Y.K., Wang F., Xue J., Zhao H., et al. Homology-independent multiallelic disruption via CRISPR/Cas9-based knock-in yields distinct functional outcomes in human cells. BMC Biol. 2018;16:151. doi: 10.1186/s12915-018-0616-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 343.Lu B., Ye J. Commentary: PROTACs make undruggable targets druggable: challenge and opportunity. Acta Pharm Sin B. 2021;11:3335–3336. doi: 10.1016/j.apsb.2021.07.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 344.He Z., Liu H., Moch H., Simon H.U. Machine learning with autophagy-related proteins for discriminating renal cell carcinoma subtypes. Sci Rep. 2020;10:720. doi: 10.1038/s41598-020-57670-y. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Acta Pharmaceutica Sinica. B are provided here courtesy of Elsevier

RESOURCES