Next Article in Journal
Rosmarinic Acid Attenuates Cadmium-Induced Nephrotoxicity via Inhibition of Oxidative Stress, Apoptosis, Inflammation and Fibrosis
Next Article in Special Issue
Membrane Lipid Composition: Effect on Membrane and Organelle Structure, Function and Compartmentalization and Therapeutic Avenues
Previous Article in Journal
Phosphoinositide 3 Kinase Signaling in Human Stem Cells from Reprogramming to Differentiation: A Tale in Cytoplasmic and Nuclear Compartments
Previous Article in Special Issue
Two Sides of the Coin: Ezrin/Radixin/Moesin and Merlin Control Membrane Structure and Contact Inhibition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Acyl-CoA-Binding Domain-Containing 3 (ACBD3; PAP7; GCP60): A Multi-Functional Membrane Domain Organizer

1
School of Life Science and Technology, ShanghaiTech University, Pudong, Shanghai 201210, China
2
School of Physical Science and Technology, ShanghaiTech University, Pudong, Shanghai 201210, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(8), 2028; https://doi.org/10.3390/ijms20082028
Submission received: 15 March 2019 / Revised: 13 April 2019 / Accepted: 15 April 2019 / Published: 24 April 2019
(This article belongs to the Special Issue Molecular Regulation of the Endomembrane System)

Abstract

:
Acyl-CoA-binding domain-containing 3 (ACBD3) is a multi-functional scaffolding protein, which has been associated with a diverse array of cellular functions, including steroidogenesis, embryogenesis, neurogenesis, Huntington’s disease (HD), membrane trafficking, and viral/bacterial proliferation in infected host cells. In this review, we aim to give a timely overview of recent findings on this protein, including its emerging role in membrane domain organization at the Golgi and the mitochondria. We hope that this review provides readers with useful insights on how ACBD3 may contribute to membrane domain organization along the secretory pathway and on the cytoplasmic surface of intracellular organelles, which influence many important physiological and pathophysiological processes in mammalian cells.
Keywords:
ACBD3; Golgi; PAP7; GCP60; PI4KB

1. Introduction

The Golgi apparatus has traditionally been considered as a central sorting station along the secretory pathway for newly synthesized secretory proteins [1,2]. More recently, however, many studies have described its roles in intracellular signaling and feedback mechanism that leads to more integrative cellular decision-making for cell division, differentiation, apoptosis, and sensing of cellular secretory activity [3,4,5].
Acyl-CoA-binding domain-containing 3 (ACBD3) was originally found in the early 2000s as a Golgi scaffolding protein that helped the function of Golgin tethers, such as Giantin and Golgin-160 in membrane trafficking and apoptosis [6,7,8,9]. ACBD3 is known to directly bind palmitoyl-CoA using its ACBD domain, while its GOLD (Golgi Dynamics) domain mediates ACBD3 targeting to the Golgi and other membranes (Figure 1) [10,11]. Over the past 20 years, highly diverse interaction partners of ACBD3 have been discovered, including PI4KB, Golgin45, TBC1D22, PPM1L, DMT1, PKA, FAPP2, TUG, PARP-1, mHtt, Numb, as well as several bacterial and viral proteins [10,12,13,14,15,16,17,18,19,20,21].
In most of these cases (except for Numb), ACBD3 seems to function as a part of a multi-protein complex and is likely to form functional membrane micro-domains at the Golgi and mitochondria. For example, ACBD3 has been shown to function as a Golgi docking site for PI4KB, which plays crucial roles for transport carrier formation at the Golgi [22]. Interestingly, ACBD3–PI4KB interaction appears to be linked to the perinuclear vacuole/sites, proximal to the Golgi, where many viruses and bacteria promote their intracellular proliferation by hijacking and exploiting membrane trafficking machineries in the host cells [12,19,22,23,24,25,26].
In this article, we discuss recent discoveries on novel ACBD3 interaction partners, through which membrane domain organization emerges as a basic theme for highly diverse physiological and pathophysiological roles played by ACBD3.

2. ACBD3 in Steroidogenesis

Liu et al. identified ACBD3 (PAP7: PBR and PKA-associated protein 7) in a yeast two-hybrid experiment, screening for PBR (peripheral-type benzodiazepine receptor) and PKA (protein kinase A)-interacting proteins [27]. PBR, a cholesterol binding protein, has been shown to mainly localize in the outer membrane of the mitochondria and participate in steroid biosynthesis by accelerating the transport of cholesterol from the outer to the inner membrane of mitochondria. PKA has also been demonstrated to be involved in steroid biosynthesis upon activation by increased concentration of cyclic AMP. PKA is a heterotetramer with two catalytic (C) subunits and two regulatory (R) subunits, which consist of four different types, namely, RIα, RIβ, RIIα, and RIIβ. PKA holoenzyme is auto-inhibitory since the C subunit is inactivated by its binding to the R subunit. Once cyclic AMP binds to the R subunit, the C subunit is released and activated. Purified recombinant GST-tagged ACBD3 fragments 228-445 (linker between the ACBP and GOLD domains plus the GOLD domain) and 212-369 (linker between the ACBP and GOLD domains plus the first half of the GOLD domain) were able to pull down PBR from mitochondrial extracts and the recombinant regulatory subunit of PKA in GST-pulldown assays, respectively.
Overexpression of the full-length ACBD3 in MA-10 Leydig cells greatly improved steroid synthesis upon stimulation by hCG, whereas lower expression of ACBD3 by antisense oligonucleotides specific to ACBD3 inhibited the steroidogenesis by MA-10 cells. Interestingly, overexpression of the fragments 228-445 and 212-369 of ACBD3, which contain the PBR and PKA binding motifs respectively, substantially reduced the progesterone production stimulated by hCG [28]. These fragments may compete with the endogenous ACBD3 and inhibit PBR or PKA binding to endogenous ACBD3, therefore preventing cholesterol transport to mitochondria mediated by the PBR–PKA–ACBD3 complex.

3. ACBD3 in Lipid Metabolism and Metabolic Homoeostasis

While ACBD3 is a positive regulator of steroidogenesis, it seems to negatively regulate de novo cholesterol synthesis through inhibiting SREBP1 (sterol regulatory element-binding protein 1) [29]. Chen et al. showed that ACBD3 was a new binding partner and maturation modulator of SREBP1. ACBD3 inhibited SREBP1-sensitive promoter activity of fatty acid synthase (FASN). Moreover, ACBD3 blocked intracellular maturation of SREBP1 probably through directly binding with the lipid regulator rather than disrupting SREBP1–SCAP–Insig1 interaction. ACBD3 overexpression inhibited de novo fatty acid synthesis by regulating FASN and ACC (acetyl-CoA carboxylase). These observations extended the regulatory properties of ACBD3 to fatty acid biogenesis beyond cholesterol biogenesis and steroidogenesis [29].
Later studies by the same group have shown that ACBD3 was involved in modulating NAD+ metabolism through activating poly(ADP-ribose) polymerase 1 (PARP1) [21]. PARP1 is a well-characterized stress response protein, which mediates various DNA repair pathways and maintains genomic stability. PARP1 also plays a regulatory role in metabolism and lipid metabolism [30]. They showed that overexpressed ACBD3 significantly reduced cellular NAD+ content via enhancing PARP1’s polymerase activity and enhancing auto-modification of the enzyme in a DNA damage-independent manner. Furthermore, they found that enhanced ERK1/2 activity (extracellular signal-regulated kinase1/2) and inhibited SREBP1-controlled fatty acid biosynthesis play crucial roles in this ACBD3-regulated PARP1 activation. Importantly, oxidative stress-induced PARP1 activation is greatly attenuated by knocking down the ACBD3 gene. These findings broaden our views on the biological roles played by ACBD3 and may provide novel insights into the mechanisms underlying how lipid-binding proteins affect metabolic homoeostasis via PARP1-involved signaling pathway.

4. ACBD3 in Regulation of Ceramide and Glucosylceramide Transport

Glycosphingolipids (GSLs) are important components of the plasma membrane (PM), with key roles in cell signaling, adhesion, proliferation, and differentiation. GSLs are synthesized at the Golgi complex from glucosylceramide (GlcCer), which is synthesized from ceramide at the cytosolic leaflet of early Golgi membranes. Upon translocation to the luminal leaflet, GlcCer is galactosylated to lactosylceramide (LacCer), which can then be converted into complex GSLs in later Golgi compartments [31,32]. GlcCer can be transported through the Golgi complex via membrane trafficking and via non-vesicular transfer owing to the action of the cytosolic GlcCer transfer protein FAPP2, which fosters GSL synthesis [33,34].
PPM1L, an endoplasmic reticulum (ER)-resident transmembrane protein phosphatase, appears to be involved in the regulation of ceramide trafficking at ER–Golgi membrane contact sites [35]. ACBD3 was shown to play a crucial role in recruiting PPM1L to ER–Golgi membrane contact sites to dephosphorylate the ceramide transport protein CERT and regulate ceramide transport [17]. More recently, ACBD3 was identified as a FAPP2-interacting partner [9]. This interaction may facilitate the recruitment of FAPP2 to ER–Golgi membrane contact sites and translocate GlcCer to the lumen of ER. GlcCer in the ER is then transported to the Golgi cisternae via vesicle trafficking for GSL biosynthesis. These studies revealed critical roles for ACBD3 in regulating cellular sphingolipid metabolism. Since ACBD3 is important in maintaining the integrity of Golgi morphology, these data also establish the importance of the Golgi complex for the transfer of GlcCer and complex GSL synthesis.

5. ACBD3 in Golgi Structure Maintenance

As mentioned in the introduction, ACBD3 was originally described as a Golgi scaffolding protein, which interacts with Golgin tethers, such as Giantin [6]. Recently, through proximity-based in vivo tagging and using a proteomics approach, we identified ACBD3 as a novel binding partner of Golgin45, a medial Golgi-localized Golgi structural protein [13]. ACBD3 appears to use its GOLD domain to interact with Golgin45 and enhance Golgin45 targeting to the Golgi. Previous studies have indicated that ACBD3 interacts with at least three other membrane trafficking proteins at the Golgi, including Giantin, Golgin-160, and TBC1D22, a Rab33b-GTPase activating protein (GAP) [6,7,16].
Unexpectedly, however, Giantin and Golgin-160 were not a part of the Golgin45–ACBD3 complex, whereas TBC1D22 was strongly detected during reciprocal co-immunoprecipitation experiments. In addition, GRASP55, a well-known Golgin45-binding partner and Golgi stacking protein, was also selectively enriched in the Golgin45–ACBD3–TBC1D22 complex. Thus, this large multi-protein complex (GRASP55–Golgin45–ACBD3–TBC1D22) seems to form a membrane micro-domain between the medial-Golgi cisternae and contributes to membrane trafficking and Golgi structure maintenance. As Rab33b had been described to be a part of Rab1 (cis)–Rab33b (medial)–Rab6 (trans) cascade in the mammalian Golgi apparatus [36,37], it is intriguing to speculate that this multi-protein complex may also play an integral role in regulating membrane identity of the medial Golgi cisternae. Further study is required to probe the validity of this hypothesis.

6. ACBD3 in Iron Uptake

At the mitochondria, stimulation of NMDAR (N-methyl-D-aspartate receptor) leads to activation of nNOS (neuron nitric oxide synthase), which generates NO (nitric oxide) and activates Dextras1 (dexamethasone-induced Ras-related protein 1) by its S-nitrosylation on cysteine 11, and results in iron uptake through Dextras1′s binding proteins, PAP7 and DMT1 (the divalent metal transporter 1) [18,38]. In undifferentiated PC12 cells, Dexras1 was found to bind to PAP7 in yeast two-hybrid, GST-pulldown and co-immunoprecipitation assays.
The Dexras1-interacting site on PAP7 has been shown to be between amino acids 193-444, composed of the linker connecting the ACBP and the GOLD domains, and the first half of the GOLD domain as well. When undifferentiated PC12 cells were treated with different NO donors, that is, GSNO (S-nitrosoglutathione), SNP (sodium nitroprusside), or DETA NONOate, Dextras1 was S-nitrosylated and NTBI (non-transferrin bound iron) uptake was enhanced in a NO donor–concentration dependent manner. In mouse primary cortical neuron cells, activation of NMDAR by treatment of NMDA increased NTBI uptake, which was inhibited by preincubation of NMDAR antagonist or nNOS gene knockout [39].

7. ACBD3 Is Exploited by Viral and Bacterial Proteins to Promote Their Replication

ACBD3 was identified in a mammalian two-hybrid assay as a host protein that utilizes its C-terminal GOLD region to bind the Aichi virus nonstructural proteins 2B, 2C, and 3A [26]. Four hours after infection, ACBD3 was found to colocalize with 2B, 2C, and 3A in clusters in the cytoplasm, whereas other Golgi proteins, namely, Giantin, GM130, and TGN46, were not distributed with ACBD3.
More recently, ACBD3 has also been shown to interact with the nonstructural protein 3A protein of plus-strand RNA viruses (picornaviruses), including human rhinovirus 14, coxsachievirus B3, and poliovirus [12,19,22,25,26,40,41,42,43,44,45,46,47]. The 3A protein is required for reorganization of the Golgi and replication of virus in host cells. In 293T cells, the Golgi proteins ACBD3 and PI4KB have been demonstrated to co-immunoprecipitate with 3A proteins through their N-terminus. Knockdown with siRNA specific to GBF1 (Golgi-specific brefeldin A resistance guanine nucleotide exchange factor 1), PI4KB, or ACBD3 greatly reduced, if not completely abolished, the replication of poliovirus and Aichi virus in HeLa cells. Mutations of the PI4KB-interacting amino acids in the N-terminus of 3A decelerated virus replication by insufficient recruitment of PI4KB [19].
ACBD3 was also demonstrated to be important for bacteria replication by presenting bacterial effector proteins close to the Golgi in host cells. ACBD3 was found to interact with bacterial protein SseG and SseF from Salmonella typhimurium in yeast two-hybrid assays and by co-immunoprecipitation in HeLa cells, but the interaction between SseF and ACBD3 is dependent on SseG. When ACBD3 was knocked down by siRNA in HeLa cells, the association of wild-type Salmonella typhimurium with the Golgi network was reduced, similar to the effect in cells infected by S. typhimurium carrying a null mutation in SseG. The bacterial replication was defective in HeLa cells infected by bacteria carrying both of the non-ACBD3-interacting mutant genes, sseGS67G and sseFC109W,D129G. However, when either sseGS67G or sseFC109W,D129G was expressed individually by S. typhimurium in HeLa cells, the association of Salmonella-containing vacuoles (SCV) and the Golgi in host cells was intact.

8. ACBD3 in Intracellular Retention of GLUT4 Storage Vesicles (GSV)

Insulin-induced translocation of glucose transporter is an important mechanism that controls blood glucose level, and any disruption in this process may lead to insulin resistance and diabetes. Belman et al. have recently shown that GSV retention in unstimulated adipocytes requires ACBD3 and overexpressed ACBD3 further enhances insulin-responsive GLUT4 translocation to the plasma membranes [14]. This effect was mediated by Sirt2-dependent acetylation of TUG (Tether containing a UBX domain for GLUT4), which in turn regulates TUG interaction with Golgi associated proteins, including ACBD3, PIST, and Golgin-160. It is interesting to note that Sirt2 targets FoxO1 and PGC1-α as well as Akt, thereby potentially affecting adipocyte differentiation and insulin-dependent signaling.

9. ACBD3 in Huntington’s Disease (HD)

Huntington’s disease is a neurodegenerative disease, caused by mutations in the gene encoding the huntingtin protein (Htt), leading to polyglutamine repeats expansion and the resulting cytotoxicity [48]. Sbodio et al. reported that ACBD3 is markedly increased in HD patients and interacts with mutant Htt (mHtt) and a small GTPase, Rhes (Ras homolog enriched in striatum) at the Golgi to form ACBD3/mHtt/Rhes complex [15]. This protein complex was found to promote mHtt/Rhes-dependent cytotoxicity [49], as depletion of ACBD3 entirely abolished mHtt/Rhes cytotoxicity, while its overexpression enhanced the cytotoxicity [15].
Strikingly, the authors also found that ACBD3 level was significantly increased by stress-inducing agents for the Golgi or the ER or the mitochondria, respectively, suggesting that prolonged cellular stress and up-regulation of ACBD3 level may be one of the causes for the cytotoxicity and onset of Huntington’s disease.

10. ACBD3 in Cancer

Lastly, but importantly, it has been reported that ACBD3 is up-regulated in breast cancer and associated with advanced pathoclinical features as well as poor prognosis in breast cancer [50]. In the study by Huang et al., overexpressing ACBD3 promoted, whereas silencing ACBD3 inhibited self-renewal of breast cancer cells in vitro. Furthermore, the tumorigenicity of breast cancer cells in vivo was also found to be increased significantly.
It was shown that up-regulating ACBD3 promoted the self-renewal and tumorigenesis of breast cancer cells via activating the Wnt/beta-catenin signaling pathway [50]. These findings seem to suggest a novel ACBD3-dependent regulatory mechanism in breast cancer.
There is emerging evidence that the capacity for self-renewal is dysregulated in cancer stem cells (CSCs) and that CSCs also possess properties that make them resistant to chemotherapy and radiation [51,52]. Notably, the expression of ACBD3 was found to be 3.8-fold higher in non-responders (PD) than in responders (PR) to gefitinib in the response of lung cancer cells to gefitinib [53]. As previously reported, ACBD3 also regulates Numb signaling during asymmetric cell division in neural progenitor cell to specify cell fate [20].
It is widely accepted that asymmetric cell division is a defining characteristics of stem cells that enables them to self-renew and differentiate. Dysregulated asymmetric division of stem cells is the root cause of tumorigenesis [54]. Clearly, it is particularly important to identify the mechanisms that convert a normal mammalian stem cell into a cancer stem cell. Further investigations of the role of ACBD3 in cancer stem cell differentiation and tumorigenesis may help increase the understanding of the onset of cancer and develop innovative therapies.

11. Conclusions

ACBD3 has unusually diverse roles both in normal cell physiology and in diseases. The data covered in this review strongly suggest that a central theme encompassing these diverse activities of ACBD3 is its ability to form distinct scaffoldings on diverse membranous environments. This characteristic diversity, in turn, seems to allow various roles via its interaction with numerous adapters/interacting proteins (Figure 2, Table 1). It still remains to be determined how these distinct membrane micro-domains containing ACBD3 can function without mixing with each other. Due to the recent advancement of big data handling and super-resolution microscopy, there will be interesting opportunities to further understand ACBD3′s diverse intracellular activities with better clarity by a more integrated analysis of basic cell biology and clinical data.

Funding

This work was supported by ShanghaiTech University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Warren, G.; Malhotra, V. The organisation of the Golgi apparatus. Curr. Opin. Cell Biol. 1998, 10, 493–498. [Google Scholar] [CrossRef]
  2. Traub, L.M.; Kornfeld, S. The trans-Golgi network: A late secretory sorting station. Curr. Opin. Cell Biol. 1997, 9, 527–533. [Google Scholar] [CrossRef]
  3. Pulvirenti, T.; Giannotta, M.; Capestrano, M.; Capitani, M.; Pisanu, A.; Polishchuk, R.S.; San Pietro, E.; Beznoussenko, G.V.; Mironov, A.A.; Turacchio, G.; et al. A traffic-activated Golgi-based signalling circuit coordinates the secretory pathway. Nat. Cell Biol. 2008, 10, 912–922. [Google Scholar] [CrossRef]
  4. Machamer, C.E. The Golgi complex in stress and death. Front. Neurosci. 2015, 9, 421. [Google Scholar] [CrossRef] [PubMed]
  5. Preisinger, C.; Korner, R.; Wind, M.; Lehmann, W.D.; Kopajtich, R.; Barr, F.A. Plk1 docking to GRASP65 phosphorylated by Cdk1 suggests a mechanism for Golgi checkpoint signalling. EMBO J. 2005, 24, 753–765. [Google Scholar] [CrossRef] [PubMed]
  6. Sohda, M.; Misumi, Y.; Yamamoto, A.; Yano, A.; Nakamura, N.; Ikehara, Y. Identification and characterization of a novel Golgi protein, GCP60, that interacts with the integral membrane protein giantin. J. Biol. Chem. 2001, 276, 45298–45306. [Google Scholar] [CrossRef] [PubMed]
  7. Sbodio, J.I.; Hicks, S.W.; Simon, D.; Machamer, C.E. GCP60 preferentially interacts with a caspase-generated golgin-160 fragment. J. Biol. Chem. 2006, 281, 27924–27931. [Google Scholar] [CrossRef]
  8. Sbodio, J.I.; Machamer, C.E. Identification of a redox-sensitive cysteine in GCP60 that regulates its interaction with golgin-160. J. Biol. Chem. 2007, 282, 29874–29881. [Google Scholar] [CrossRef] [PubMed]
  9. Liao, J.; Guan, Y.; Chen, W.; Shi, C.; Yao, D.; Wang, F.; Lam, S.M.; Shui, G.; Cao, X. ACBD3 is required for FAPP2 transferring glucosylceramide through maintaining the Golgi integrity. J. Mol. Cell Biol. 2018. [Google Scholar] [CrossRef]
  10. Fan, J.; Liu, J.; Culty, M.; Papadopoulos, V. Acyl-coenzyme A binding domain containing 3 (ACBD3; PAP7; GCP60): An emerging signaling molecule. Prog. Lipid Res. 2010, 49, 218–234. [Google Scholar] [CrossRef] [PubMed]
  11. Soupene, E.; Kuypers, F.A. Ligand binding to the ACBD6 protein regulates the acyl-CoA transferase reactions in membranes. J. Lipid Res. 2015, 56, 1961–1971. [Google Scholar] [CrossRef]
  12. Ishikawa-Sasaki, K.; Sasaki, J.; Taniguchi, K. A complex comprising phosphatidylinositol 4-kinase IIIbeta, ACBD3, and Aichi virus proteins enhances phosphatidylinositol 4-phosphate synthesis and is critical for formation of the viral replication complex. J. Virol. 2014, 88, 6586–6598. [Google Scholar] [CrossRef]
  13. Yue, X.; Bao, M.; Christiano, R.; Li, S.; Mei, J.; Zhu, L.; Mao, F.; Yue, Q.; Zhang, P.; Jing, S.; et al. ACBD3 functions as a scaffold to organize the Golgi stacking proteins and a Rab33b-GAP. FEBS Lett. 2017, 591, 2793–2802. [Google Scholar] [CrossRef]
  14. Belman, J.P.; Bian, R.R.; Habtemichael, E.N.; Li, D.T.; Jurczak, M.J.; Alcazar-Roman, A.; McNally, L.J.; Shulman, G.I.; Bogan, J.S. Acetylation of TUG protein promotes the accumulation of GLUT4 glucose transporters in an insulin-responsive intracellular compartment. J. Biol. Chem. 2015, 290, 4447–4463. [Google Scholar] [CrossRef]
  15. Sbodio, J.I.; Paul, B.D.; Machamer, C.E.; Snyder, S.H. Golgi protein ACBD3 mediates neurotoxicity associated with Huntington’s disease. Cell Rep. 2013, 4, 890–897. [Google Scholar] [CrossRef]
  16. Greninger, A.L.; Knudsen, G.M.; Betegon, M.; Burlingame, A.L.; DeRisi, J.L. ACBD3 interaction with TBC1 domain 22 protein is differentially affected by enteroviral and kobuviral 3A protein binding. MBio 2013, 4, e00098-13. [Google Scholar] [CrossRef]
  17. Shinoda, Y.; Fujita, K.; Saito, S.; Matsui, H.; Kanto, Y.; Nagaura, Y.; Fukunaga, K.; Tamura, S.; Kobayashi, T. Acyl-CoA binding domain containing 3 (ACBD3) recruits the protein phosphatase PPM1L to ER-Golgi membrane contact sites. FEBS Lett. 2012, 586, 3024–3029. [Google Scholar] [CrossRef]
  18. Okazaki, Y.; Ma, Y.; Yeh, M.; Yin, H.; Li, Z.; Yeh, K.Y.; Glass, J. DMT1 (IRE) expression in intestinal and erythroid cells is regulated by peripheral benzodiazepine receptor-associated protein 7. Am. J. Physiol. 2012, 302, G1180–G1190. [Google Scholar] [CrossRef]
  19. Greninger, A.L.; Knudsen, G.M.; Betegon, M.; Burlingame, A.L.; Derisi, J.L. The 3A protein from multiple picornaviruses utilizes the golgi adaptor protein ACBD3 to recruit PI4KIIIbeta. J. Virol. 2012, 86, 3605–3616. [Google Scholar] [CrossRef]
  20. Zhou, Y.; Atkins, J.B.; Rompani, S.B.; Bancescu, D.L.; Petersen, P.H.; Tang, H.; Zou, K.; Stewart, S.B.; Zhong, W. The mammalian Golgi regulates numb signaling in asymmetric cell division by releasing ACBD3 during mitosis. Cell 2007, 129, 163–178. [Google Scholar] [CrossRef]
  21. Chen, Y.; Bang, S.; Park, S.; Shi, H.; Kim, S.F. Acyl-CoA-binding domain containing 3 modulates NAD+ metabolism through activating poly(ADP-ribose) polymerase 1. Biochem. J. 2015, 469, 189–198. [Google Scholar] [CrossRef] [PubMed]
  22. Klima, M.; Toth, D.J.; Hexnerova, R.; Baumlova, A.; Chalupska, D.; Tykvart, J.; Rezabkova, L.; Sengupta, N.; Man, P.; Dubankova, A.; et al. Structural insights and in vitro reconstitution of membrane targeting and activation of human PI4KB by the ACBD3 protein. Sci. Rep. 2016, 6, 23641. [Google Scholar] [CrossRef] [PubMed]
  23. Yu, X.J.; Liu, M.; Holden, D.W. Salmonella Effectors SseF and SseG Interact with Mammalian Protein ACBD3 (GCP60) To Anchor Salmonella-Containing Vacuoles at the Golgi Network. MBio 2016, 7, e00474-16. [Google Scholar] [CrossRef] [PubMed]
  24. Greninger, A.L. Picornavirus--host interactions to construct viral secretory membranes. Prog. Mol. Biol. Transl. Sci. 2015, 129, 189–212. [Google Scholar]
  25. Teoule, F.; Brisac, C.; Pelletier, I.; Vidalain, P.O.; Jegouic, S.; Mirabelli, C.; Bessaud, M.; Combelas, N.; Autret, A.; Tangy, F.; et al. The Golgi protein ACBD3, an interactor for poliovirus protein 3A, modulates poliovirus replication. J. Virol. 2013, 87, 11031–11046. [Google Scholar]
  26. Sasaki, J.; Ishikawa, K.; Arita, M.; Taniguchi, K. ACBD3-mediated recruitment of PI4KB to picornavirus RNA replication sites. Embo J. 2012, 31, 754–766. [Google Scholar] [CrossRef]
  27. Liu, J.; Matyakhina, L.; Han, Z.; Sandrini, F.; Bei, T.; Stratakis, C.A.; Papadopoulos, V. Molecular cloning, chromosomal localization of human peripheral-type benzodiazepine receptor and PKA regulatory subunit type 1A (PRKAR1A)-associated protein PAP7, and studies in PRKAR1A mutant cells and tissues. FASEB J. 2003, 17, 1189–1191. [Google Scholar] [CrossRef]
  28. Liu, J.; Rone, M.B.; Papadopoulos, V. Protein-protein interactions mediate mitochondrial cholesterol transport and steroid biosynthesis. J. Biol. Chem. 2006, 281, 38879–38893. [Google Scholar] [CrossRef]
  29. Chen, Y.; Patel, V.; Bang, S.; Cohen, N.; Millar, J.; Kim, S.F. Maturation and activity of sterol regulatory element binding protein 1 is inhibited by acyl-CoA binding domain containing 3. PLoS ONE 2012, 7, e49906. [Google Scholar] [CrossRef]
  30. Bai, P. Biology of Poly(ADP-Ribose) Polymerases: The Factotums of Cell Maintenance. Mol. Cell 2015, 58, 947–958. [Google Scholar] [CrossRef]
  31. Hannun, Y.A.; Obeid, L.M. Principles of bioactive lipid signalling: Lessons from sphingolipids. Nat. Rev. Mol. Cell Biol. 2008, 9, 139–150. [Google Scholar] [CrossRef]
  32. Hannun, Y.A.; Obeid, L.M. Sphingolipids and their metabolism in physiology and disease. Nat. Rev. Mol. Cell Biol. 2018, 19, 175–191. [Google Scholar] [CrossRef]
  33. D’Angelo, G.; Polishchuk, E.; Di Tullio, G.; Santoro, M.; Di Campli, A.; Godi, A.; West, G.; Bielawski, J.; Chuang, C.C.; van der Spoel, A.C.; et al. Glycosphingolipid synthesis requires FAPP2 transfer of glucosylceramide. Nature 2007, 449, 62–67. [Google Scholar] [CrossRef]
  34. D’Angelo, G.; Uemura, T.; Chuang, C.C.; Polishchuk, E.; Santoro, M.; Ohvo-Rekila, H.; Sato, T.; Di Tullio, G.; Varriale, A.; D’Auria, S.; et al. Vesicular and non-vesicular transport feed distinct glycosylation pathways in the Golgi. Nature 2013, 501, 116–120. [Google Scholar] [CrossRef]
  35. Saito, S.; Matsui, H.; Kawano, M.; Kumagai, K.; Tomishige, N.; Hanada, K.; Echigo, S.; Tamura, S.; Kobayashi, T. Protein phosphatase 2Cepsilon is an endoplasmic reticulum integral membrane protein that dephosphorylates the ceramide transport protein CERT to enhance its association with organelle membranes. J. Biol. Chem. 2008, 283, 6584–6593. [Google Scholar] [CrossRef]
  36. Pusapati, G.V.; Luchetti, G.; Pfeffer, S.R. Ric1-Rgp1 complex is a guanine nucleotide exchange factor for the late Golgi Rab6A GTPase and an effector of the medial Golgi Rab33B GTPase. J. Biol. Chem. 2012, 287, 42129–42137. [Google Scholar] [CrossRef]
  37. Pfeffer, S.R. Rab GTPase localization and Rab cascades in Golgi transport. Biochem. Soc. Trans. 2012, 40, 1373–1377. [Google Scholar] [CrossRef]
  38. Chen, Y.; Mathias, L.; Falero-Perez, J.M.; Kim, S.F. PKA-mediated phosphorylation of Dexras1 suppresses iron trafficking by inhibiting S-nitrosylation. FEBS Lett. 2015, 589, 3212–3219. [Google Scholar] [CrossRef]
  39. Wilkinson, G.; Dennis, D.; Schuurmans, C. Proneural genes in neocortical development. Neuroscience 2013, 253, 256–273. [Google Scholar] [CrossRef]
  40. Delang, L.; Paeshuyse, J.; Neyts, J. The role of phosphatidylinositol 4-kinases and phosphatidylinositol 4-phosphate during viral replication. Biochem. Pharmacol. 2012, 84, 1400–1408. [Google Scholar] [CrossRef]
  41. Arita, M.; Wakita, T.; Shimizu, H. Valosin-containing protein (VCP/p97) is required for poliovirus replication and is involved in cellular protein secretion pathway in poliovirus infection. J. Virol. 2012, 86, 5541–5553. [Google Scholar] [CrossRef]
  42. Ishikawa-Sasaki, K.; Nagashima, S.; Taniguchi, K.; Sasaki, J. Model of OSBP-Mediated Cholesterol Supply to Aichi Virus RNA Replication Sites Involving Protein-Protein Interactions among Viral Proteins, ACBD3, OSBP, VAP-A/B, and SAC1. J. Virol. 2018, 92, e01952-17. [Google Scholar] [CrossRef]
  43. Xiao, X.; Lei, X.; Zhang, Z.; Ma, Y.; Qi, J.; Wu, C.; Xiao, Y.; Li, L.; He, B.; Wang, J. Enterovirus 3A facilitates viral replication by promoting PI4KB-ACBD3 interaction. J. Virol. 2017. [Google Scholar] [CrossRef]
  44. Roberts, M.F. Defining a Viral Membrane-Remodeling Complex on an Atomic Level. Structure 2017, 25, 3–4. [Google Scholar] [CrossRef]
  45. McPhail, J.A.; Ottosen, E.H.; Jenkins, M.L.; Burke, J.E. The Molecular Basis of Aichi Virus 3A Protein Activation of Phosphatidylinositol 4 Kinase IIIbeta, PI4KB, through ACBD3. Structure 2017, 25, 121–131. [Google Scholar] [CrossRef]
  46. Klima, M.; Chalupska, D.; Rozycki, B.; Humpolickova, J.; Rezabkova, L.; Silhan, J.; Baumlova, A.; Dubankova, A.; Boura, E. Kobuviral Non-structural 3A Proteins Act as Molecular Harnesses to Hijack the Host ACBD3 Protein. Structure 2017, 25, 219–230. [Google Scholar] [CrossRef]
  47. Dorobantu, C.M.; van der Schaar, H.M.; Ford, L.A.; Strating, J.R.; Ulferts, R.; Fang, Y.; Belov, G.; van Kuppeveld, F.J. Recruitment of PI4KIIIbeta to coxsackievirus B3 replication organelles is independent of ACBD3, GBF1, and Arf1. J. Virol. 2014, 88, 2725–2736. [Google Scholar] [CrossRef]
  48. MacDonald, M.E.; Ambrose, C.M.; Duyao, M.P.; Myers, R.H.; Lin, C.; Srinidhi, L.; Barnes, G.; Taylor, S.A.; James, M.; Groot, N.; et al. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. Cell 1993, 72, 971–983. [Google Scholar] [CrossRef]
  49. Subramaniam, S.; Sixt, K.M.; Barrow, R.; Snyder, S.H. Rhes, a striatal specific protein, mediates mutant-huntingtin cytotoxicity. Science 2009, 324, 1327–1330. [Google Scholar] [CrossRef]
  50. Huang, Y.; Yang, L.; Pei, Y.Y.; Wang, J.; Wu, H.; Yuan, J.; Wang, L. Overexpressed ACBD3 has prognostic value in human breast cancer and promotes the self-renewal potential of breast cancer cells by activating the Wnt/beta-catenin signaling pathway. Exp. Cell Res. 2018, 363, 39–47. [Google Scholar] [CrossRef]
  51. Lytle, N.K.; Barber, A.G.; Reya, T. Stem cell fate in cancer growth, progression and therapy resistance. Nat. Rev. Cancer 2018, 18, 669–680. [Google Scholar] [CrossRef]
  52. O’Brien, C.A.; Kreso, A.; Ryan, P.; Hermans, K.G.; Gibson, L.; Wang, Y.; Tsatsanis, A.; Gallinger, S.; Dick, J.E. ID1 and ID3 regulate the self-renewal capacity of human colon cancer-initiating cells through p21. Cancer Cell 2012, 21, 777–792. [Google Scholar] [CrossRef]
  53. Kakiuchi, S.; Daigo, Y.; Ishikawa, N.; Furukawa, C.; Tsunoda, T.; Yano, S.; Nakagawa, K.; Tsuruo, T.; Kohno, N.; Fukuoka, M.; et al. Prediction of sensitivity of advanced non-small cell lung cancers to gefitinib (Iressa, ZD1839). Hum. Mol. Genet. 2004, 13, 3029–3043. [Google Scholar] [CrossRef]
  54. Knoblich, J.A. Asymmetric cell division: Recent developments and their implications for tumour biology. Nat. Rev. Mol. Cell Biol. 2010, 11, 849–860. [Google Scholar] [CrossRef]
Figure 1. (A) Schematic representation of the acyl-CoA-binding domain-containing 3 (ACBD3) domain structure with known binding sites. (ACBD: acyl-CoA binding domain; CAR: charged amino acid region; Q: glutamine-rich domain) (B) Pseudoatomic model of ACBD3, based on previously published GOLD (PDB code: 5TDQ) and Q domain (PDB code: 2N72) structures plus a homologous ACBD domain crystal structure from bovine ACBD1 (PDB code: 1ACA), which binds palmitoyl-CoA directly [10,22]. Crystal structure of full length ACBD3 and its ACBD domain are currently not available.
Figure 1. (A) Schematic representation of the acyl-CoA-binding domain-containing 3 (ACBD3) domain structure with known binding sites. (ACBD: acyl-CoA binding domain; CAR: charged amino acid region; Q: glutamine-rich domain) (B) Pseudoatomic model of ACBD3, based on previously published GOLD (PDB code: 5TDQ) and Q domain (PDB code: 2N72) structures plus a homologous ACBD domain crystal structure from bovine ACBD1 (PDB code: 1ACA), which binds palmitoyl-CoA directly [10,22]. Crystal structure of full length ACBD3 and its ACBD domain are currently not available.
Ijms 20 02028 g001
Figure 2. Schematic illustration of diverse roles of ACBD3 and its interaction partners at the Golgi, mitochondria, and cytoplasm in the cells. ACBD3′s physiological roles range from steroidogenesis, neurogenesis, lipid metabolism, cell division, apoptosis, membrane trafficking to proposed roles in various diseases like cancer, diabetes, viral and bacterial replication in host cells, and Huntington’s disease (HD), etc.
Figure 2. Schematic illustration of diverse roles of ACBD3 and its interaction partners at the Golgi, mitochondria, and cytoplasm in the cells. ACBD3′s physiological roles range from steroidogenesis, neurogenesis, lipid metabolism, cell division, apoptosis, membrane trafficking to proposed roles in various diseases like cancer, diabetes, viral and bacterial replication in host cells, and Huntington’s disease (HD), etc.
Ijms 20 02028 g002
Table 1. Previously identified ACBD3-interacting proteins.
Table 1. Previously identified ACBD3-interacting proteins.
Protein NamesMolecular WeightIntracellular LocalizationFunctionsReferences
Giantin372 kDaGolgiGolgi organization; ER to Golgi transport[6]
Golgin160160 kDaGolgiMaintenance of Golgi structure[7,8]
Golgin4545 kDaGolgi and nucleusRequired for normal Golgi structure and for protein transport from the ER through the Golgi to the cell surface[11]
Numb72 kDaplasma membrane, nucleus and cytosolPlays a role in the determination of cell fates during development[19]
PI4KB92 kDaGolgi, endosome and cytosolMay regulate Golgi disintegration/reorganization during mitosis; Involved in Golgi-to-plasma membrane trafficking[10,17,21,42,44,45]
TUG60 kDaplasma membrane, nucleus and cytosolTethering protein that sequesters GLUT4-containing vesicles in the cytoplasm in the absence of insulin. Modulates the amount of GLUT4 that is available at the cell surface. Enhances VCP methylation catalyzed by VCPKMT.[12]
Htt35 kDacytoskeleton, nucleus, cytosol, ER, Golgi and endosomeHuntingtin is a disease gene linked to Huntington’s disease; May play a role in microtubule-mediated transport or vesicle function[13]
PBR18 kDamitochondriaPromotes the transport of cholesterol across mitochondrial membranes and may play a role in lipid metabolism[26]
PRKAR1A43 kDaplasma membrane and cytosolRegulatory subunit of the cAMP-dependent protein kinases involved in cAMP signaling in cells[26]
3A protein Nonstructural viral
protein required for RNA replication
[14,23,24,42,44,45]
SseF and SseG Both are (Salmonella pathogenicity island 2) SPI-2-encoded effectors which are necessary for the retention of tightly clustered bacterial microcolonies in close proximity to the MTOC and the Golgi[22]
DMT162 kDaplasma membrane, mitochondria, nucleus lysosome, endosome and GolgiImportant in metal transport, in particular iron[16]
PARP1110 kDaMitochondria and nucleusInvolved in the base excision repair (BER) pathway, by catalyzing the poly(ADP-ribosyl)ation of a limited number of acceptor proteins involved in chromatin architecture and in DNA metabolism[20]
SREBP1120 kDaER, nucleus, cytosol and GolgiTranscriptional activator required for lipid homeostasis[28]
FAPP258 kDanucleus and GolgiCargo transport protein that is required for apical transport from the Golgi complex; Mediates the non-vesicular transport of glucosylceramide (GlcCer) from the trans-Golgi network (TGN) to the plasma membrane and plays a pivotal role in the synthesis of complex glycosphingolipids[9]
TBC1D22A47 kDaPerinuclear, GolgiGTPase activating protein (GAP) for Rab33b; membrane trafficking[14]
PPM1L41 kDaERHelps regulate ceramide transport from the ER to the Golgi; Acts as a suppressor of the SAPK signaling pathways[15]

Share and Cite

MDPI and ACS Style

Yue, X.; Qian, Y.; Gim, B.; Lee, I. Acyl-CoA-Binding Domain-Containing 3 (ACBD3; PAP7; GCP60): A Multi-Functional Membrane Domain Organizer. Int. J. Mol. Sci. 2019, 20, 2028. https://doi.org/10.3390/ijms20082028

AMA Style

Yue X, Qian Y, Gim B, Lee I. Acyl-CoA-Binding Domain-Containing 3 (ACBD3; PAP7; GCP60): A Multi-Functional Membrane Domain Organizer. International Journal of Molecular Sciences. 2019; 20(8):2028. https://doi.org/10.3390/ijms20082028

Chicago/Turabian Style

Yue, Xihua, Yi Qian, Bopil Gim, and Intaek Lee. 2019. "Acyl-CoA-Binding Domain-Containing 3 (ACBD3; PAP7; GCP60): A Multi-Functional Membrane Domain Organizer" International Journal of Molecular Sciences 20, no. 8: 2028. https://doi.org/10.3390/ijms20082028

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

Article Metrics

Back to TopTop