Open Access

The role of the Golgi apparatus in disease (Review)

  • Authors:
    • Jianyang Liu
    • Yan Huang
    • Ting Li
    • Zheng Jiang
    • Liuwang Zeng
    • Zhiping Hu
  • View Affiliations

  • Published online on: February 4, 2021     https://doi.org/10.3892/ijmm.2021.4871
  • Article Number: 38
  • Copyright: © Liu et al. This is an open access article distributed under the terms of Creative Commons Attribution License [CC BY_NC 4.0].

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

The Golgi apparatus is known to underpin many important cellular homeostatic functions, including trafficking, sorting and modifications of proteins or lipids. These functions are dysregulated in neurodegenerative diseases, cancer, infectious diseases and cardiovascular diseases, and the number of disease‑related genes associated with Golgi apparatus is on the increase. Recently, many studies have suggested that the mutations in the genes encoding Golgi resident proteins can trigger the occurrence of diseases. By summarizing the pathogenesis of these genetic diseases, it was found that most of these diseases have defects in membrane trafficking. Such defects typically result in mislocalization of proteins, impaired glycosylation of proteins, and the accumulation of undegraded proteins. In the present review, we aim to understand the patterns of mutations in the genes encoding Golgi resident proteins and decipher the interplay between Golgi resident proteins and membrane trafficking pathway in cells. Furthermore, the detection of Golgi resident protein in human serum samples has the potential to be used as a diagnostic tool for diseases, and its central role in membrane trafficking pathways provides possible targets for disease therapy. Thus, we also introduced the clinical value of Golgi apparatus in the present review.

1. Introduction

The Golgi apparatus is a processing and sorting hub in the transport and targeting of soluble cargo proteins and lipids to different destinations in the cell (1). Considering its central role in the secretory pathway, alterations in the structure and function of the Golgi apparatus are expected to affect the homeostasis of cellular proteins and lipids. Increasing evidence suggests that structural changes and functional disorder of the Golgi apparatus are involved in many human diseases such as neurodegenerative diseases (2-4), ischemic stroke (5,6), cardiovascular diseases (7,8), pulmonary arterial hypertension (9,10), infectious diseases (11-13), and cancer (14). However, much work is still needed to elucidate how the Golgi apparatus affects the progression of these diseases.

In this review, we describe the central roles of the Golgi apparatus in cells, and discuss diseases associated with structural changes and functional disorder of the Golgi apparatus. We highlight some of the studies that explore links between mutation in genes encoding Golgi resident proteins and human diseases. By analyzing their pathophysiology, we found that the majority of genes leading to human diseases are involved in membrane trafficking. Considering the mechanistic links between Golgi resident proteins, membrane trafficking, and the development of genetic diseases, we suggest a term for these disorders based on their similar pathophysiology: Golgi apparatus membrane trafficking disorders.

2. Golgi apparatus structure and function

In 1898, the Italian anatomist Camillio Golgi initially described the cell organelle that bears his name, the Golgi apparatus (15). The Golgi apparatus is characterized by a series of flattened, cisternal membrane structures forming the so-called Golgi stack, which is surrounded by vesicles. Based on the distribution of resident proteins, the Golgi stack can be divided into three regions: The cis-, medial-, and trans-Golgi cisternae (16). The Golgi stacks in vertebrate cells are laterally interconnected by tubular membranes and exhibit a twisted ribbon-like network known as the Golgi ribbon (17). The structure of the Golgi ribbon is supported by the Golgi matrix (18). The Golgi matrix is believed to comprise highly dynamic structural proteins, which is important for structural integrity and vesicular trafficking.

The Golgi apparatus has two main functions. The first is the post-translational protein modification. Similar to glycosylation, it is a common post-translational modification occurring in the endoplasmic reticulum (ER) and Golgi and the glycan processing occurs throughout the Golgi stacks. The second is the sorting, packing, routing and recycling of these modified cargos to the appropriate cellular destinations (1). The main secretory pathway can be divided into the following steps (19): First, newly synthesized proteins or lipids enter the exit sites of the ER and are sorted into budding vesicles that are dependent on the COPII. Second, vesicles move to the ER-Golgi intermediate compartment (ERGIC) and forward to the cis-Golgi networks (CGN). Third, proteins or lipids enter cis-Golgi cisternae and move towards the trans-Golgi cisternae. Vesicular transport and cisternal maturation are the two classical models of intra-Golgi transport (20). The vesicular transport model proposes that Golgi cisternae are static, and the cargos are transported through them by COPI vesicles. The cisternal maturation model suggests that cisternae are dynamic structures, while Golgi enzymes are recycled via retrograde transport of COPI vesicles. Fourth, vesicles reach the trans-Golgi networks (TGN), which are involved in the sorting of products to their final destinations such as lysosomes, endosomes, or the plasma membrane.

3. Structural and functional changes of the Golgi apparatus in diseases

The structural integrity of the Golgi apparatus is vital for its normal function, and Golgi fragmentation could result in a wide range of diseases and disorders. Functional changes of the Golgi Apparatus include perturbations in Golgi pH, aberrant Golgi glycosylation, and membrane trafficking. Golgi fragmentation has been found to often be an early causative event in the process of cell apoptosis (21,22). With pharmacological or oxidative stress, a series of changes occur in the Golgi apparatus, such as cargo overloading, ionic imbalance, and abnormal luminal acidity. These changes can lead to defects in membrane trafficking. We previously presented 'Golgi stress' as a new concept to explain the Golgi-specific stress response (23). The Golgi stress response constitutes autoregulation to repair the Golgi apparatus and may initiate signaling pathways to alleviate stress. The nucleus signaling pathways of the Golgi stress response was identified in a previous study: The procaspase-2/golgin-160, TFE3, HSP47, and the CREB3-ARF4 pathways (24). If these pathways fail to repair overstimulation, the Golgi is completely disassembled, inducing cell apoptosis.

Apoptosis triggered by structural changes and functional disorder of the Golgi contributes to the pathogenesis of many diseases, such as neurodegenerative diseases (25), ischemic stroke (5,6), cardiovascular diseases (26), pulmonary arterial hypertension (9,10), infectious diseases (12,13), and cancer (27). A summary of diseases relating to the Golgi apparatus, classified on the basis of the main organ affected is shown in Fig. 1.

Neurodegenerative disease

Structural and functional changes of the Golgi apparatus are associated with several neurodegenerative diseases, such as Amyotrophic lateral sclerosis (28), Alzheimer's disease (29), Parkinson's disease (3), Huntington's disease (30), Creutzfeldt-Jacob disease (31) and multiple system atrophy (32). Golgi fragmentation is not a consequence of apoptosis, but a very early event in the pathological cascade in neurodegenerative disorders and precedes other pathological changes in the neuron (33). Golgi fragmentation may alter neuronal physiology, and induce failures in transport to axons, dendrites, and synapses (34). Finally, Golgi alteration may trigger a stress response and, as consequence, result in neuronal death. Furthermore, Golgi fragmentation in neurodegenerative disease alters protein trafficking and production, such as amyloid precursor protein in Alzheimer's disease (35), and sodium-dependent vitamin C transporter 2 in Huntington's disease (36). The causes of Golgi fragmentation in neurodegenerative diseases may be diverse. First, alteration of the microtubule and microfilament stabilization may also be the cause (37). In Alzheimer's disease and other tauopathies, tau-induced microtubule-bundling may result in Golgi fragmentation (38). Furthermore, perturbations in Golgi pH are also responsible for Golgi fragmentation. The Purkinje cells from the Golgi pH regulator conditional knockout mice exhibited Golgi fragmentation, followed by axonal degeneration and neuronal loss (39).

Infectious disease

Golgi fragmentation has been identified in diseases such as infection by Orf virus (12), Chlamydia trachomatis (40,41), Hepatitis C virus (HCV) (42), Human Rhinovirus (HRV) (13), and Rickettsia rickettsii (43). Golgi fragmentation in these infectious diseases is mainly reflected in two aspects: i) Escaping from the immune response. In infected cells, Golgi fragmentation reduces MHC class I complex surface expression by defective membrane trafficking (43,44), which may aid in escaping host cellular immune recognition (12); ii) Enhancing viral replication. In human rhinovirus-1A infection, the Golgi in host cells is fragmented and rearranged into vesicles that appear to be used as the membrane source for the assembly of viruses (45). Similarly, in Oropouche virus replication, proteins in the endosomal sorting complex required for transport in the host cell are hijacked in Golgi cisternae to mediate remodeling of Golgi membranes, resulting in enlargement of the Golgi stacks, where the endosomal sorting complex required for transport participates in the assembly of viral factories (46). Thus, structural changes in the Golgi apparatus may enhance viral replication in infectious diseases by providing membranes.

Cancer

Aberrant Golgi glycosylation is reported to regulate invasion of cancer cells, such as in prostate (47), breast (48), and gastric cancer (49). Golgi glycosylation is involved in basic molecular and cellular biology processes occurring in cancer, such as cell signaling transduction and communication, cancer cell dissociation and invasion, cell-matrix adhesion, cancer angiogenesis, immune regulation and metastasis (50). Similar to epithelial cadherin, a transmembrane glycoprotein, is involved in epithelial cell-cell adhesion in tumors (51). The Golgi glycosylation of N-linked glycans on epithelial cadherin can affect the epithelial-mesenchymal transition, which is related to the formation of metastatic lesions (49). This process is suggested to help cancer cells leave their original position during wound healing and other normal physiological processes, which is an essential mechanism for metastasis and diffusion of cancer cells (52,53). The GOLPH3 complex is an important molecular component in the process of Golgi-driven tumor progression. The role of the GOLPH3 complex in cancer includes: i) Regulating Golgi glycosylation, which is important in driving the cancer phenotype (54); ii) promoting the cellular DNA damage response that enhances cellular survival under DNA damage (55); iii) interacting with components of the retromer complex that enhances growth-factor-induced mTOR signaling (56); and iv) regulating cell migration by promoting reorientation of the Golgi apparatus towards the leading edge (57). In addition to GOLPH3, the Golgi protein GM130 is important in Golgi glycosylation and protein membrane trafficking in cancer cells. Downregulation of GM130 induces autophagy, inhibits glycosylation, decreases angiogenesis, and suppresses tumorigenesis (58). In general, aberrant Golgi glycosylation causes carcinogenesis, but may also be a consequence of cancer progression.

Other diseases

Golgi dysfunction was also observed in pulmonary arterial hypertension, and cardiovascular diseases. In an in vivo model of pulmonary arterial hypertension, Golgi dysfunction and intracellular trafficking with trapping of diverse vesicle tethers, giantin, p115, and soluble N-ethylmaleimide-sensitive factor attachment protein receptors (SNAREs) were observed in the Golgi membranes of enlarged pulmonary arterial endothelial cells and smooth muscle cells (9,10,59). Golgi-mediated membrane trafficking dysfunctions play important roles in the pathogenesis of pulmonary arterial hypertension (60).

Structural changes and functional disorder of the Golgi apparatus have been identified in many cardiovascular diseases, such as heart failure, dilated cardiomyopathy, arrhythmia, and chronic arial fibrillation (61-64). A previous review clarified the relationship between the Golgi apparatus and various cardiovascular diseases (26). For example, in dilated cardiomyopathy patients, morphological changes in Golgi vesicle are consistent with the secretion of natriuretic peptide as the rate of protein secretion affects the morphology and size of Golgi vesicles (7). In addition, the Golgi vesicle area is inversely proportional to the left ventricular end-diastolic diameter and the end-systolic diameter, and is proportional to the left ventricular ejection fraction (65).

4. Mutant Golgi resident proteins involved in disease

In addition to being an intermediate site in pathogenic cascades in diseases, the Golgi apparatus can be the primary target for diseases caused by genetic mutations in Golgi resident proteins. Mutations in proteins localized to the Golgi apparatus can be deleterious for the structure and function of this organelle, impeding membrane trafficking pathways through it (Fig. 2) and resulting in disease. We highlight some of the studies that explore links between Golgi resident proteins and disease.

Golgi matrix protein and diseases

Adjacent Golgi stacks are linked by tubules forming a membrane network termed the Golgi ribbon (66). This structure is a highly ordered and continuous structure that is adjacent to the nucleus. The Golgi ribbon comprises proteins that mediate cisternal stacking and the material supporting the Golgi ribbon is the Golgi matrix (67). The concept of the Golgi matrix was introduced by Slusarewicz and colleagues, who isolated a detergent-insoluble, salt-resistant Golgi fraction in 1994 (18). The main function of the Golgi matrix is maintaining normal structure and mediating protein trafficking through the Golgi cisternae. During cisternal progression, the Golgi matrix must be dynamic to adapt to Golgi structural changes.

Golgi matrix proteins include golgins and Golgi reassembly stacking proteins (GRASPs) (67), both of which are important for maintaining Golgi structure and regulating protein and lipid trafficking through the stacks. Golgins are a family of conserved coiled-coil proteins that were originally identified as a group of Golgi-localized antigens (68,69). The golgins not only capture incoming vesicles, but also clearly distinguish vesicles from different origins (70). GRASPs include GRASP65 (71) and GRASP55 (72). The former localizes to the cis-Golgi cisternae while the latter localizes to the medial/trans-Golgi cisternae. The functions of GRASPs include Golgi structure formation, specific cargo transport, apoptosis, and cell migration (73).

Given the important multiple functions of Golgi matrix proteins, mutation of Golgi matrix proteins has serious consequences on health. Increasing studies support that the mutation of Golgi matrix proteins including GM130, Bicaudal-D (BICD), GMAP-210, giantin (74), and SCYL1BP1 (also known as GORAB) (75), leads to diseases. The present review included some proteins as examples to elaborate on the pathogenic mechanism of Golgi matrix proteins.

The first example is GM130 (also known as GOLGA2), the first identified Golgi matrix protein (76). GM130 is a peripheral membrane protein attached to the Golgi membrane that is important in maintaining the adaxial Golgi reticular structure (77). In neurodegenerative diseases, GM130 knockout in hippocampal neurons is reported to cause damage to dendritic structures (78). In mouse neuron experiments, specific knockout of GM130 resulted in disruption of the Golgi architecture and positioning in cerebellar Purkinje cells and to deficient secretory cargo trafficking. As a consequence, progressive cerebellar atrophy of Purkinje cells resulted in delayed movement and ataxia in mice (79). This animal experimental study indicates that GM130 mutations are causative in neurodegenerative disease.

A second example is BICD, a golgin that interacts with Rab6 on the TGN (80). Of two homologous sequences, BICD1 and BICD2, the latter binds to a subgroup of motility protein activator proteins and is a connecting molecule between the motility protein and cargo (81). High expression of BICD in normal nervous systems is important for maintaining the normal lamellar structure of the cerebral cortex, hippocampus, and cerebellar cortex (82). The brain cortex, hippocampus and cerebellar cortex neurons of BICD2-knockout mice have impaired migration function (82,83) and eventually, damage the brain and cerebellar cortex layer structure. Previous findings showed that, missense mutations in BICD resulted in spinal muscular atrophy (84,85) and hereditary spastic paraplegia (86) by changing the normal morphological structure of the golgi. The core pathogenetic mechanism may be a BICD2 mutation resulting in abnormal cargo trafficking in motor neurons. This trafficking results in neuronal growth disorders and eventually neuronal dysfunction.

The third example is giantin, encoded by the Golgb1 gene. Giantin is a member of the golgin family and is a tethering factor for COPI vesicles and functions in the CGN (87). Mutations in the Golgb1 gene lead to lack of expression of giantin protein and a pleiotropic phenotype including osteochondrodysplasia in a rat model (88) and a ciliopathy-like phenotype in a zebrafish model (74). Both pathogenetic mechanisms involve disturbance of extracellular matrix components, which are transported by intracellular membrane trafficking systems. Giantin knockout leads to changes in expression of Golgi-resident glycosyltransferases, which could affect extracellular matrix deposition (89).

The fourth example is GORAB (also known as SCYL1BP1). GORAB, localized to the trans-side of the Golgi, is a member of the golgin family and interacts with Rab6. Mutation in GORAB results in gerodermia osteodysplastica (GO) characterized by wrinkly skin and osteoporosis (75). GORAB functions in COPI trafficking, and acts as a scaffolding factor for COPI assembly at the TGN by interacting with Scyl1. GORAB mutations perturb COPI assembly at the TGN, and result in reduced recycling of COPI-mediated retrieval of trans-Golgi enzymes and improper glycosylation (90).

A final example of the effects of loss of expression of a Golgi matrix protein is GMAP-210 (also known as TRIP11). This CGN golgin acts in asymmetric membrane tethering (91). In animal experiments, a nonsense mutation in Trip11 led to a loss of GMAP-210, which led to abnormal Golgi-mediated glycosylation and cellular transport of proteins in chondrocytes and osteoblasts of mice (92). Similarly, GMAP-210 mutations were found in patients with human chondrodysplasia achondrogenesis 1A (92), and odontochondrodysplasia (93).

Other Golgi resident proteins and diseases

In addition to matrix proteins, several proteins that localize to Golgi membranes are also important for normal Golgi structure and function such as the tethering factors Rab GTPases and SNAREs, which regulate the specific targeting and fusion of transport carriers with Golgi membranes. The maintenance of Golgi luminal ion concentrations depends on the secretory pathway Ca2+/Mn2+ ATPases and vacuolar H+ ATPase (V-ATPase). Therefore, the impaired performance of mutated Golgi resident proteins creates serious and highly diverse pathologies in the Golgi. Emerging studies on patient genetics have identified mutations in Golgi resident protein-coding genes that are related to diseases. We focus on some of these proteins, and discuss the activities of mutated Golgi resident proteins that result in disease.

Golgi ion pump

The release and uptake of Ca2+ by Golgi membranes is mainly mediated by secretory pathway Ca2+/Mn2+ ATPases (SPCA1 and SPCA2), which are encoded by the ATP2C1/ATP2C2 genes. The proteins transfer Ca2+ from the cytoplasm to the Golgi and maintain the stability of intracellular free Ca2+ (94). The maintenance of Golgi luminal Ca2+ and Mn2+ directly affects the optimal activity of Golgi glycosyltransferase and the trafficking of cell adhesion proteins to the cell plasma membrane (95). Knockdown of SPCA1 affects the morphology and structure of the Golgi and causes mis-localization of proteins. Clinically, mutations in the ATP2C1 gene on chromosome 3q21 can lead to Hailey-Hailey disease, an autosomal dominant skin disorder in humans (96,97). The possible pathogenetic mechanism may be dysfunction in Ca2+ signaling at the Golgi membrane and dysfunction of processing, modification and trafficking of desmosomal proteins (98).

Golgi acidity is an important role for maintaining the morphological integrity of the Golgi and transporting various kinds of cargo (99,100). Under normal conditions, the Golgi cavity is weakly acidic and the pH of the Golgi reticular structure decreases gradually from the CGN to the TGN (101). The Golgi luminal pH is regulated by V-ATPase (102), AE2a HCO3-/Cl- exchanger, and Golgi pH regulator (103). Luminal pH is closely tied to Golgi function. Partial V-ATPase dysfunction is related to multiple disease states (104). ATP6V1E1, ATP6V1A, and ATP6V0A2 encode different subunits of the V-ATPase pump. A study showed that Golgi subunit-isoform of the V-ATPase (ATP6V0A2) mutations lead to structural changes in the extracellular matrix that is responsible for skin elasticity (105). Clinically, the dysfunction of the Golgi-localized V-ATPase caused by mutations in the ATP6VOA2 gene is directly related to cutis laxa. Mutations in ATP6V1E1 or ATP6V1A also cause autosomal-recessive cutis laxa (106). Autosomal recessive cutis laxa type II is a heterogeneous condition characterized by sagging, inelastic, and wrinkled skin (107,108). The mechanism may involve impaired intracellular acidification of the Golgi and damaged retrograde trafficking from the Golgi to the ER (100,108).

ATP7A and ATP7B are the key regulators of cellular Cu2+ metabolism. Under basal conditions (normal copper levels), ATP7A is located in the TGN and travels to the plasma membrane at high copper levels. Mutations in the ATP7A result in mislocalization of ATP7A protein and impaired copper-responsive trafficking between the TGN and plasma membrane, which contributes to the development of Menkes disease (109). Menkes disease is a lethal multisystemic disorder characterized by neurodegeneration and connective tissue abnormalities as well as typical sparse and steely hair. Similarly, mutations in the ATP7B contributes to the development of Wilson's disease (110). Wilson's disease, also known as hepatolenticular degeneration, results in hepatic and/or neurological deficits, including dystonia and parkinsonism.

Golgi resident glycosyltransferase

The Golgi apparatus is an important organelle for the post-translational modification of cargos. The post-translational modification of secreted and membrane proteins is mediated by the Golgi resident enzymes such as glycosyltransferases, glycosidases, and kinases. Glycosylation is an enzymatic reaction that chemically links monosaccharides or polysaccharides (glycans) to other saccharides, proteins, or lipids (111). Golgi glycosylation is a modification by Golgi-resident glycosylation enzymes including glycosidases and glycosyltransferases (112). The normal function of Golgi glycosylation depends on the precise Golgi localization and normal activities of Golgi resident enzymes. The proper localization of Golgi resident enzymes is controlled by finely regulated vesicular trafficking in the Golgi. If the balance between anterograde and retrograde trafficking is defective, Golgi glycosylation is affected, resulting in Golgi glycosylation abnormalities (113). Mutations in Golgi resident putative glycosyltransferases are directly linked to human congenital muscular dystrophies: Like-acetylglucosaminyl-transferase (LARGE) in congenital muscular dystrophy syndrome (114), fukutin in Fukuyama-type congenital muscular dystrophy (115), and fukutin-related protein in band muscular dystrophy syndrome (116). These mutations appear to affect cell migration in the developing brain, resulting in combined clinical manifestations in muscle and brain development. In an animal model, mutations in Golgi resident glycosyltransferases are also associated with the neurodegenerative disease, such as ST3GAL5,β1,4-gala ctosyltransferase 4 (B4GalT4) (117), and glycosyltransferase 8 domain containing 1 (GLT8D1). GLT8D1 is a glycosyltransferase enzyme located in the Golgi apparatus. A recent study reported that mutated GLT8D1 induces motor deficits in zebrafish embryos consistent with amyotrophic lateral scle- rosis (118). However, another study suggested that GLT8D1 is not likely the causative gene for ALS in mainland China (119).

Rab GTPase

Rab proteins are members of the small Ras-like GTPase family that regulate the four steps of membrane transport by recruiting effector molecules. Golgi-associated Rab proteins including Rab1, Rab2, Rab6, Rab18, Rab33B, and Rab43 have a central role in Golgi organization and membrane trafficking (120). Rab33B is localized to medial-Golgi cisternae and is important in Golgi-to-ER retrograde trafficking. Rab39B, a neuronal-specific protein, is a novel Rab GTPase that localizes to the Golgi and is related to synapse formation. Mutations in the Rab33B coding gene cause Smith-McCort dysplasia (121) and mutations in the Rab39B gene cause X-linked mental retardation (122).

SNAREs

SNAREs are proteins involved in docking and fusion of transport to intermediate membranes. Golgi SNAP receptor complex member 2 (GOSR2) is a member of the SNAREs family that localizes to the CGN and is involved in ER-to-Golgi trafficking (123). Homozygous mutations in GOSR2 lead to progressive myoclonus epilepsy (124). Clinical manifestations include early ataxia, myoclonus, and convulsive seizures. A possible mechanism involves GOSR2 mutations leading to GOSR2 protein that cannot be localized to the CGN and blocks SNAREs complex formation. SNAREs complex dysfunction could lead to the impaired fusion of vesicles with cis-Golgi cisternae, hindering ER-to-Golgi membrane trafficking. The perturbation of early ER-to-Golgi transport may result in changes in the regulated release of neurotransmitters and proper sorting of neurotransmitter receptors at synapses in neurons, potentially leading to epilepsy (125,126).

5. Golgi apparatus membrane trafficking disorders

In the above section, we introduced the pathophysiology of some diseases related to Golgi resident proteins. A summary of genetic diseases caused by mutations in genes encoding Golgi resident proteins is presented in Table I. By analyzing the pathophysiology of these diseases, we found that the majority of genes leading to human diseases are involved in defects in membrane trafficking (Fig. 2). For example, TRAPPC2 mutation, involving the membrane trafficking pathway between ER-to-Golgi in bone cells and chondrocytes, results in X-linked spondyloepiphyseal dysplasia tarda (127). The conserved oligomeric Golgi (COG) complex is a conserved, hetero-octameric protein complex localized in the Golgi cis/medial cisternae (128). In addition to the COG3 subunit, mutations in seven other COG subunits result in human congenital disorders of glycosylation (CD G) type II, which is mainly marked by misregulation of protein glycosylation, and defects in retrograde trafficking through the Golgi (129,130). The mutation in FGD1 resulting in Aarskog-Scott syndrome may lead to the obstruction of post-Golgi trafficking, such as the Golgi-to-plasma membrane trafficking pathway (131). Mutation in TRIP11 mainly involves ER to ERGIC and anterograde trafficking (132). Therefore, membrane trafficking defects play a major role in the pathogenic process of mutation in genes encoding Golgi resident protein. Intracellular membrane trafficking is a fundamental process responsible for compartmentalization of the biosynthesis pathway and secretion cargos, including hormones, growth factors, antibodies, matrix and serum proteins, digestive enzymes, and many more. Defective membrane trafficking results in protein sorting defects, undegraded proteins due to defective Golgi-to-lysosome trafficking, downregulation of protein secretion, and mislocalization of proteins.

Table I

Human diseases caused by mutations in genes encoding Golgi resident proteins.

Table I

Human diseases caused by mutations in genes encoding Golgi resident proteins.

GenefunctionDiseaseMain clinical manifestationCellular effect(Refs.)
AP1S2Coat adapterX-linked mental retardation syndromeMental retardationBrain-specific defect of AP-1-dependent intracellular protein trafficking(165)
AP3D1Coat adapterHermansky-Pudlak syndromeImmunodeficiency; Neurodevelopmental delay; SeizureImpaired lysosomal trafficking(166)
ARFGEF2GTPase activatorPeriventricular nodular heterotopiaMalformation of cortical developmentDefective TGN-cell membrane trafficking(167)
ATP2C1Ion pumpHailey-Hailey diseaseSkin disorderDefective trafficking of desmosomal proteins to cell membrane(96)
ATP6V1AIon pumpCutis laxa type IIWrinkled skinDefective retrograde transport; Abnormal glycosylation(106,168)
ATP6V1E1Ion pumpCutis laxa type IIWrinkled skinDefective retrograde transport; Abnormal glycosylation(106)
ATP6VOA2Ion pumpCutis laxa type IIWrinkled skinDefective Golgi trafficking; Abnormal glycosylation of CDG-II(108)
ATP7AIon pumpMenkes disease; Occipital horn diseaseNeurodegeneration; Connective tissue disorderDefective Golgi trafficking of copper;(109)
ATP7BIon pumpWilson's diseaseHepatic and/or neurological disorderDefective Golgi trafficking of copper(110)
ATXN2SignalingSpinocerebellar ataxia type 2Progressive ataxia; slow saccadesDisrupted calcium homeostasis(169)
Bicaudal-DGolginSMA; HSP Neurodegenerationdefective targeting and transport of Golgi resident proteins.(84,86)
COGTetheringCDG-type IINeurodegenerative disorderDefective retrograde and endosome-to-TGN trafficking; Abnormal glycosylation(170)
COPACoatCOPA syndromeInterstitial lung, joint and kidney disorderDefective membrane trafficking(171)
DENND5AGTPase activatorEpileptic EncephalopathyRefractory seizures and cognitive arrestDefective endosome-TGN trafficking(172)
DYMUnknown Dyggve-Melchior-Clausen syndrome Spondyloepimetaphyseal dysplasia; intellectual disabilityDefective ER-Golgi trafficking(173)
FGD1GTPase activatorAarskog-Scott syndromeFaciogenital dysplasiaReduction in FGD1 trafficking from Golgi(131)
FKRP GlycosyltransferasesLimb girdle muscular dystrophyMuscular dystrophyAbnormal glycosylation(116)
Fukutin GlycosyltransferasesFCMDMuscular dystrophyAbnormal glycosylation; Impaired ER-to-Golgi trafficking of mutant protein(115)
GOSR2SNAREProgressive myoclonus epilepsySeizureMislocalization of mutant protein to cis-Golgi; Defective cis to trans Golgi compartment trafficking(124)
HERC1GTPase activatorIdiopathic intellectual disabilityIntellectual disabilityMisregulation of mTOR pathway(174)
LARGE GlycosyltransferasesCongenital muscular dystrophy Type 1DMuscular dystrophyAbnormal glycosylation(114)
OSBPL2Lipid transportAutosomal dominant nonsyndromic hearing lossHearing lossAbnormal lipid metabolism(175)
RAB33BRab GTPaseSmith-McCort dysplasiaSkeletal dysplasiaGolgi fragmentation; Defective Golgi membrane trafficking(176)
RAB39BRab GTPaseX-linked Mental retardationMental Retardation; Autism; Epilepsy; MacrocephalyDefective Golgi membrane trafficking(122)
S1PSerine protease Spondyloepimetaphyseal dysplasiaSkeletal dysplasiaDefective Golgi-to-lysosome transport(177,178)
SCYL1BP1GolginGerodermia osteodysplasticaOsteoporosis; Wrinkly skinReduced recycling of trans-Golgi enzymes; Defective COPI traffic and glycosylation(75,90)
SLC35A1CMP Synal TransporterCDG-IINeurodegenerative disorderAbnormal glycosylation(179)
SLC35A2UDP Gal TransporterCDGDevelopmental delay; Seizures; AtaxiaAbnormal glycosylation(180)
TMEM165Ion pumpCDG-IINeurodegenerative disorderMislocalization of mutant protein resulting in abnormal Golgi glycosylation(134)
TRAPPC11TetheringCongenital muscular dystrophyMuscular dystrophyDefective trafficking and Hypoglycosylation of mutant protein(135)
TRAPPC2TetheringSpondyloepiphyseal dysplasia tardaSkeletal dysplasiaAbnormal trafficking between ER and Golgi(136)
TRIP11GolginAchondrogenesis type 1A; OdontochondrodysplasiaSkeletal dysplasiaGolgi fragmentation; Abnormal Golgi-mediated glycosylation(92,93)
VPS53UnknownProgressive cerebello-cerebral atrophy type 2Mental retardation; Microcephaly; EpilepsyImpaired NPC2 protein sorting to lysosome and cholesterol accumulation(181)

Considering the mechanistic links between Golgi resident proteins, membrane trafficking, and the development of genetic diseases, we suggest a term for these disorders based on their similar pathophysiology: Golgi apparatus membrane trafficking disorders. It is a group of genetic diseases in which the mutation of the gene encoding Golgi resident protein results in membrane trafficking defects within the cells. Golgi apparatus membrane trafficking defects typically result in the accumulation of undegraded proteins, mislocalization of proteins, and impaired glycosylation of proteins. However, the cascade events following the Golgi apparatus and defective membrane trafficking, ultimately leading to human diseases, remain to be clarified in further research.

Although the Golgi apparatus-mediated membrane trafficking pathway exists in all kinds of tissues and organs in human, the trafficking defects on tissues is often selective. The most sensitive to membrane trafficking defects is the nervous system, skin, bone, cartilage, and skeletal muscle and the reasons for mutations occurring in these genes mostly affecting these tissues remain to be elucidated. Firstly, neurons are extraordinarily polarized cells, the extension of dendrites and axons requires a significant expansion of the cell surface area, and new plasma membrane proteins must be delivered through the membrane trafficking. For the nervous system, intracellular trafficking functionally impacts neuronal development, homeostasis, as well as neurodegeneration (133). Secondly, it is generally known that skin, bone, cartilage, and skeletal muscle fiber comprise large amounts of the extracellular matrix which define the structure and physical properties. Almost all extracellular matrix components are transported by intra- cellular trafficking systems. Alterations in Golgi apparatus membrane trafficking can lead to glycosylation abnormalities. The assembly and maintenance of the extracellular matrix are susceptible to impairment of matrix protein glycosylation. Thus, the skin, bone, cartilage, and skeletal muscle are most sensitive to impaired glycosylation of cargo proteins, and membrane trafficking defects. Therefore, the loss of some Golgi resident proteins, such as ATP6V1A, ATP6V1E1 (106), ATP6VOA2 (108), TMEM165 (134), GOLGB1 (88), SCYL1BP1 (75), TRAPPC11 (135), TRAPPC2 (136), and TRIP11 (92), manifest primarily in these matrix-rich tissues.

6. Clinical value of Golgi apparatus

The Golgi apparatus participates in the occurrence and development of disease and could be the key to finding new targets for disease diagnosis and therapy.

Biomarker discovery

Golgi glycoprotein 73 (GP73, also referred to as GOLPH2), a resident Golgi membrane protein, is predominantly expressed in biliary epithelial cells in the normal human liver (137). GP73 expression is upregulated in chronic Hepatitis B virus (HBV) infection (138), chronic HCV infection (139), non-alcoholic fatty liver disease (140), and hepatocellular carcinoma (HCC) (141,142). Serum GP73, a new marker for HCC, is reported to appear earlier than serum α-fetoprotein. The combined detection of serum α-fetoprotein and GP73 can improve sensitivity and specificity for HCC diagnosis (143,144). However, several studies showed GP73 levels were not higher in HCC patients than in patients with other liver diseases such as cirrhosis (145,146). In addition to being a marker, the expression of GP73 is critical for chemo- therapeutic resistance in HCC cell lines (147).

Transmembrane protein 165 (TMEM165) functions in ion homeostasis, membrane trafficking, and glycosylation in the Golgi apparatus (148). Findings of a study showed that mutations in TMEM165 cause CDG type II in humans (134). Other research has found that expression of TMEM165 mRNA and protein is apparently increased in HCC patient tissues and contributes to the invasive activity of cancer cells (149). This result indicates that TMEM165 is a possible biomarker for HCC. GS28 is a member of the SNAREs protein family. GS28 protein immunoreactivity was observed in both nuclear and cytoplasmic compartments of cancer cells. High nuclear expression of GS28 is associated with poor prognosis for colorectal (150) and cervical cancer patients (151).

Anti-Golgi antibodies (AGAs) were first found in 1982 in the serum of patients with Sjogren's syndrome complicated with lymphoma (152). AGAs have also been found in other immunological diseases (153-155). Currently, at least 20 Golgi autoantigens are known, including golgin-97, golgin-67, golgin-245, golgin-95, golgin-160, and giantin. AGA positivity is commonly found in connective tissue diseases such as Sjogren's syndrome, rheumatoid arthritis, and systemic lupus erythematosus (154,156); cerebellar malignant disease such as idiopathic late-onset cerebellar ataxia (157); infectious diseases such as HBV/HCV infection, Epstein-Barr virus infection and HIV infection (155,158,159); and tumors, such as HCC and lung cancer (160). Although AGAs are not specific to any disease, their clinical detection may be helpful for classifying and following the progress of some connective tissue diseases. For example, compared to anti-BICD2-negative patients, single specificity anti-BICD2 patients may be more associated with inflammatory myopathy and interstitial lung disease (161).

Biomarkers are crucial for early diagnosis, assessing response to treatment, and classifying diseases into subtypes. Biomarker discovery involves many critical steps such as clinical study design, sample collection, data integration, and protein/peptide identification and preservation. These steps should be carefully controlled before confirmation and verification. Therefore, in clinical applications, these biomarkers are potential diagnostic markers. Large-scale investigations are needed and more sensitive and specific detection methods need to be researched.

Golgi-based therapeutics

In addition to biomarker discovery, the functions of the Golgi apparatus and its associated molecules in maintaining cell structural integrity and its central role in membrane trafficking pathways provide possible targets for disease therapy. These targets may be direct, due to genetic disease (Table I), or indirect, as in cancer. Compared to non-transformed and normal cells, cancer cells have morphological and functional changes in the Golgi apparatus that drive invasion and migration in a unique microenvironment. These changes provide therapeutic targets for interventions. A research team developed a bovine serum albumin pH-responsive photothermal ablation agent that preferentially accumulates in the Golgi of cancer cells compared to normal cells due to morphological changes in the Golgi apparatus (162). The agent is activated by the weakly acidic microenvironment of the Golgi in cancer cells for photothermal therapy. In this method, a photothermal ablation agent converts light energy into heat and kills cancer cells with high specificity and minimal invasiveness by hyper-pyrexia (162). Another research team developed a prodrug nanoparticle system, which appeared to target the Golgi apparatus and realized retinoic acid release under an acidic environment. The retinoic acid-conjugated chondroitin sulfate could reduce the expression of metastasis-associated proteins by inducing Golgi fragmentation (163). Those findings suggest that the Golgi apparatus is a promising target for the development of novel drugs. A review summarized small molecules as drugs targeting the Golgi apparatus for the treatment of diseases (164), such as LTX-401, inhibitors of Golgi-associated lipid transfer proteins, glucosylceramide synthase inhibitors, O-glycosylation inhibitors, PI4KIIIb inhibitors and inhibitors of ARF activation. Whether these drugs that target the Golgi apparatus can be applied in clinical practice needs to be determined.

7. Conclusion

The central role of the Golgi apparatus in critical cell processes such as the transport, processing, and sorting of proteins and lipids has placed it at the forefront of cell science. Several previous studies have suggested that the Golgi apparatus plays a critical role in diseases, particularly in neurodegenerative diseases. However, few studies focus on human diseases caused by mutations in genes encoding Golgi resident proteins and summarize the common features of these genetic diseases. In the present review, we summed up the genetic diseases caused by mutations in genes encoding Golgi resident proteins. By analyzing their pathophysiology, we identified that the majority of genes are involved in membrane trafficking. The nervous system, skin, bone, cartilage, and skeletal muscle are the most sensitive tissues to defective membrane trafficking. It is reasonable to hope that our basic knowledge of Golgi-mediated membrane trafficking will continue to provide insights into the pathogenesis of genetic diseases and that studies of these diseases will continue to enhance our under- standing of the critical role of the Golgi apparatus in diseases. In addition, the finding of Golgi-related biomarker and Golgi-based therapeutics further emphasize the importance of Golgi apparatus in human pathology. Taken together, advances in Golgi apparatus biology provide opportunities to translate discoveries into clinical medicine. Thus, we highlighted the importance of underlying clinical insights and provided a new direction for future research.

Funding

The present study was supported by grants from the National Natural Science Foundation of China (grant no. 81974213).

Availability of data and materials

Not applicable.

Authors' contributions

JL and YH were mainly responsible for collecting relevant information and completing this review. ZJ, LZ and TL were mainly responsible for consulting literature materials and revising the manuscript. ZH was responsible for the conception of this review and the assignment of tasks. There was no additional assistance with manuscript preparation. All authors read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Acknowledgments

Not applicable.

References

1 

Rios RM and Bornens M: The Golgi apparatus at the cell centre. Curr Opin Cell Biol. 15:60–66. 2003. View Article : Google Scholar : PubMed/NCBI

2 

Gautam M, Jara JH, Sekerkova G, Yasvoina MV, Martina M and Özdinler PH: Absence of alsin function leads to corticospinal motor neuron vulnerability via novel disease mechanisms. Hum Mol Genet. 25:1074–1087. 2016. View Article : Google Scholar : PubMed/NCBI

3 

Rendón WO, Martínez-Alonso E, Tomás M, Martínez-Martínez N and Martínez-Menárguez JA: Golgi fragmentation is Rab and SNARE dependent in cellular models of Parkinson's disease. Histochem Cell Biol. 139:671–684. 2013. View Article : Google Scholar

4 

Brandstaetter H, Kruppa AJ and Buss F: Huntingtin is required for ER-to-Golgi transport and for secretory vesicle fusion at the plasma membrane. Dis Model Mech. 7:1335–1340. 2014. View Article : Google Scholar : PubMed/NCBI

5 

Yuan D, Liu C and Hu B: Dysfunction of membrane trafficking leads to ischemia-reperfusion injury after transient cerebral ischemia. Transl Stroke Res. 9:215–222. 2018. View Article : Google Scholar :

6 

Li T, You H, Mo X, He W, Tang X, Jiang Z, Chen S, Chen Y, Zhang J and Hu Z: GOLPH3 mediated Golgi stress response in modulating N2A cell death upon oxygen-glucose deprivation and reoxygenation injury. Mol Neurobiol. 53:1377–1385. 2016. View Article : Google Scholar

7 

Tarazón E, Roselló-Lletí E, Ortega A, Gil-Cayuela C, González-Juanatey JR, Lago F, Martínez-Dolz L, Portolés M and Rivera M: Changes in human Golgi apparatus reflect new left ventricular dimensions and function in dilated cardiomyopathy patients. Eur J Heart Fail. 19:280–282. 2017. View Article : Google Scholar

8 

Stancu CS, Toma L and Sima AV: Dual role of lipoproteins in endothelial cell dysfunction in atherosclerosis. Cell Tissue Res. 349:433–446. 2012. View Article : Google Scholar : PubMed/NCBI

9 

Lee J, Reich R, Xu F and Sehgal PB: Golgi, trafficking, and mitosis dysfunctions in pulmonary arterial endothelial cells exposed to monocrotaline pyrrole and NO scavenging. Am J Physiol Lung Cell Mol Physiol. 297:L715–L728. 2009. View Article : Google Scholar : PubMed/NCBI

10 

Sehgal PB, Mukhopadhyay S, Xu F, Patel K and Shah M: Dysfunction of Golgi tethers, SNAREs, and SNAPs in monocrotaline-induced pulmonary hypertension. Am J Physiol Lung Cell Mol Physiol. 292:L1526–L1542. 2007. View Article : Google Scholar : PubMed/NCBI

11 

Zhu H, Li H, Wang P, Chen M, Huang Z, Li K, Li Y, He J, Han J and Zhang Q: Persistent and acute chlamydial infections induce different structural changes in the Golgi apparatus. Int J Med Microbiol. 304:577–585. 2014. View Article : Google Scholar : PubMed/NCBI

12 

Rohde J, Emschermann F, Knittler MR and Rziha HJ: Orf virus interferes with MHC class I surface expression by targeting vesicular transport and Golgi. BMC Vet Res. 8:1142012. View Article : Google Scholar : PubMed/NCBI

13 

Mousnier A, Swieboda D, Pinto A, Guedán A, Rogers AV, Walton R, Johnston SL and Solari R: Human rhinovirus 16 causes Golgi apparatus fragmentation without blocking protein secretion. J Virol. 88:11671–11685. 2014. View Article : Google Scholar : PubMed/NCBI

14 

Tan X, Banerjee P, Guo HF, Ireland S, Pankova D, Ahn YH, Nikolaidis IM, Liu X, Zhao Y, Xue Y, et al: Epithelial-to- mesenchymal transition drives a pro-metastatic Golgi compaction process through scaffolding protein PAQR11. J Clin Invest. 127:117–131. 2017. View Article : Google Scholar

15 

Golgi C: On the structure of nerve cells. 1898. J Microsc. 155:3–7. 1989. View Article : Google Scholar : PubMed/NCBI

16 

Mollenhauer HH and Morré DJ: Perspectives on Golgi apparatus form and function. J Electron Microsc Tech. 17:2–14. 1991. View Article : Google Scholar : PubMed/NCBI

17 

Storrie B, White J, Röttger S, Stelzer EH, Suganuma T and Nilsson T: Recycling of golgi-resident glycosyltransferases through the ER reveals a novel pathway and provides an explanation for nocodazole-induced Golgi scattering. J Cell Biol. 143:1505–1521. 1998. View Article : Google Scholar : PubMed/NCBI

18 

Slusarewicz P, Nilsson T, Hui N, Watson R and Warren G: Isolation of a matrix that binds medial Golgi enzymes. J Cell Biol. 124:405–413. 1994. View Article : Google Scholar : PubMed/NCBI

19 

Papanikou E and Glick BS: Golgi compartmentation and identity. Curr Opin Cell Biol. 29:74–81. 2014. View Article : Google Scholar : PubMed/NCBI

20 

Glick BS and Luini A: Models for Golgi traffic: A critical assessment. Cold Spring Harb Perspect Biol. 3:a0052152011. View Article : Google Scholar : PubMed/NCBI

21 

Sundaramoorthy V, Sultana JM and Atkin JD: Golgi fragmentation in amyotrophic lateral sclerosis, an overview of possible triggers and consequences. Front Neurosci. 9:4002015. View Article : Google Scholar : PubMed/NCBI

22 

Gonatas NK, Stieber A and Gonatas JO: Fragmentation of the Golgi apparatus in neurodegenerative diseases and cell death. J Neurol Sci. 246:21–30. 2006. View Article : Google Scholar : PubMed/NCBI

23 

Jiang Z, Hu Z, Zeng L, Lu W, Zhang H, Li T and Xiao H: The role of the Golgi apparatus in oxidative stress: Is this organelle less significant than mitochondria? Free Radic Biol Med. 50:907–917. 2011. View Article : Google Scholar : PubMed/NCBI

24 

Liu JY, He JL, Huang Y, Xiao H, Jiang Z and Hu ZP: The Golgi apparatus in neurorestoration. J Neuroresstoratology. 7:116–128. 2019. View Article : Google Scholar

25 

Fan J, Hu Z, Zeng L, Lu W, Tang X, Zhang J and Li T: Golgi apparatus and neurodegenerative diseases. Int J Dev Neurosci. 26:523–534. 2008. View Article : Google Scholar : PubMed/NCBI

26 

Lu L, Zhou Q, Chen Z and Chen L: The significant role of the Golgi apparatus in cardiovascular diseases. J Cell Physiol. 233:2911–2919. 2018. View Article : Google Scholar

27 

Millarte V and Farhan H: The Golgi in cell migration: Regulation by signal transduction and its implications for cancer cell metastasis. ScientificWorldJournal. 2012:4982782012. View Article : Google Scholar : PubMed/NCBI

28 

Mourelatos Z, Gonatas NK, Stieber A, Gurney ME and Dal Canto MC: The Golgi apparatus of spinal cord motor neurons in transgenic mice expressing mutant Cu, Zn superoxide dismutase becomes fragmented in early, preclinical stages of the disease. Proc Natl Acad Sci USA. 93:5472–5477. 1996. View Article : Google Scholar

29 

Joshi G, Bekier ME II and Wang Y: Golgi fragmentation in Alzheimer's disease. Front Neurosci. 9:3402015. View Article : Google Scholar : PubMed/NCBI

30 

Strehlow AN, Li JZ and Myers RM: Wild-type huntingtin participates in protein trafficking between the Golgi and the extracellular space. Hum Mol Genet. 16:391–409. 2007. View Article : Google Scholar

31 

Sakurai A, Okamoto K, Fujita Y, Nakazato Y, Wakabayashi K, Takahashi H and Gonatas NK: Fragmentation of the Golgi apparatus of the ballooned neurons in patients with corticobasal degeneration and Creutzfeldt-Jakob disease. Acta Neuropathol. 100:270–274. 2000. View Article : Google Scholar : PubMed/NCBI

32 

Sakurai A, Okamoto K, Yaguchi M, Fujita Y, Mizuno Y, Nakazato Y and Gonatas NK: Pathology of the inferior olivary nucleus in patients with multiple system atrophy. Acta Neuropathol. 103:550–554. 2002. View Article : Google Scholar : PubMed/NCBI

33 

van Dis V, Kuijpers M, Haasdijk ED, Teuling E, Oakes SA, Hoogenraad CC and Jaarsma D: Golgi fragmentation precedes neuromuscular denervation and is associated with endosome abnormalities in SOD1-ALS mouse motor neurons. Acta Neuropathol Commun. 2:382014. View Article : Google Scholar : PubMed/NCBI

34 

Pottorf T, Mann A, Fross S, Mansel C and Vohra BPS: Nicotinamide mononucleotide adenylyltransferase 2 maintains neuronal structural integrity through the maintenance of golgi structure. Neurochem Int. 121:86–97. 2018. View Article : Google Scholar : PubMed/NCBI

35 

Joshi G and Wang Y: Golgi defects enhance APP amyloidogenic processing in Alzheimer's disease. Bioessays. 37:240–247. 2015. View Article : Google Scholar

36 

Covarrubias-Pinto A, Parra AV, Mayorga-Weber G, Papic E, Vicencio I, Ehrenfeld P, Rivera FJ and Castro MA: Impaired intracellular trafficking of sodium-dependent vitamin C transporter 2 contributes to the redox imbalance in Huntington's disease. J Neurosci Res. 99:223–235. 2021. View Article : Google Scholar

37 

Mani M, Thao DT, Kim BC, Lee UH, Kim DJ, Jang SH, Back SH, Lee BJ, Cho WJ, Han IS and Park JW: DRG2 knock- down induces Golgi fragmentation via GSK3β phosphorylation and microtubule stabilization. Biochim Biophys Acta Mol Cell Res. 1866:1463–1474. 2019. View Article : Google Scholar : PubMed/NCBI

38 

Rodríguez-Cruz F, Torres-Cruz FM, Monroy-Ramírez HC, Escobar-Herrera J, Basurto-Islas G, Avila J and García-Sierra F: Fragmentation of the Golgi apparatus in neuroblastoma cells is associated with tau-induced ring-shaped microtubule bundles. J Alzheimers Dis. 65:1185–1207. 2018. View Article : Google Scholar : PubMed/NCBI

39 

Sou YS, Kakuta S, Kamikubo Y, Niisato K, Sakurai T, Parajuli LK, Tanida I, Saito H, Suzuki N, Sakimura K, et al: Cerebellar neurodegeneration and neuronal circuit remodeling in Golgi pH regulator-deficient mice. eNeuro. 6:ENEURO.0427-18.2019. 2019. View Article : Google Scholar : PubMed/NCBI

40 

Heuer D, Rejman Lipinski A, Machuy N, Karlas A, Wehrens A, Siedler F, Brinkmann V and Meyer TF: Chlamydia causes fragmentation of the Golgi compartment to ensure reproduction. Nature. 457:731–735. 2009. View Article : Google Scholar

41 

Pruneda JN, Bastidas RJ, Bertsoulaki E, Swatek KN, Santhanam B, Clague MJ, Valdivia RH, Urbé S and Komander D: A chlamydia effector combining deubiquitination and acetylation activities induces Golgi fragmentation. Nat Microbiol. 3:1377–1384. 2018. View Article : Google Scholar : PubMed/NCBI

42 

Hansen MD, Johnsen IB, Stiberg KA, Sherstova T, Wakita T, Richard GM, Kandasamy RK, Meurs EF and Anthonsen MW: Hepatitis C virus triggers Golgi fragmentation and autophagy through the immunity-related GTPase M. Proc Natl Acad Sci USA. 114:E3462–E3471. 2017. View Article : Google Scholar : PubMed/NCBI

43 

Aistleitner K, Clark T, Dooley C and Hackstadt T: Selective frag- mentation of the trans-Golgi apparatus by Rickettsia rickettsii. PLoS Pathog. 16:e10085822020. View Article : Google Scholar

44 

Ganesan M, Mathews S, Makarov E, Petrosyan A, Kharbanda KK, Kidambi S, Poluektova LY, Casey CA and Osna NA: Acetaldehyde suppresses HBV-MHC class I complex presentation on hepatocytes via induction of ER stress and Golgi fragmentation. Am J Physiol Gastrointest Liver Physiol. 319:G432–G442. 2020. View Article : Google Scholar : PubMed/NCBI

45 

Quiner CA and Jackson WT: Fragmentation of the Golgi apparatus provides replication membranes for human rhinovirus 1A. Virology. 407:185–195. 2010. View Article : Google Scholar : PubMed/NCBI

46 

Barbosa NS, Mendonça LR, Dias MVS, Pontelli MC, da Silva EZM, Criado MF, da Silva-Januário ME, Schindler M, Jamur MC, Oliver C, et al: ESCRT machinery components are required for orthobunyavirus particle production in Golgi compartments. PLoS Pathog. 14:e10070472018. View Article : Google Scholar : PubMed/NCBI

47 

Petrosyan A, Holzapfel MS, Muirhead DE and Cheng PW: Restoration of compact Golgi morphology in advanced prostate cancer enhances susceptibility to galectin-1-induced apoptosis by modifying mucin O-glycan synthesis. Mol Cancer Res. 12:1704–1716. 2014. View Article : Google Scholar : PubMed/NCBI

48 

Tokuda E, Itoh T, Hasegawa J, Ijuin T, Takeuchi Y, Irino Y, Fukumoto M and Takenawa T: Phosphatidylinositol 4-phosphate in the Golgi apparatus regulates cell-cell adhesion and invasive cell migration in human breast cancer. Cancer Res. 74:3054–3066. 2014. View Article : Google Scholar : PubMed/NCBI

49 

Zhao J, Yang C, Guo S and Wu Y: GM130 regulates epithelial-to-mesenchymal transition and invasion of gastric cancer cells via snail. Int J Clin Exp Pathol. 8:10784–10791. 2015.PubMed/NCBI

50 

Pinho SS and Reis CA: Glycosylation in cancer: Mechanisms and clinical implications. Nat Rev Cancer. 15:540–555. 2015. View Article : Google Scholar : PubMed/NCBI

51 

Pinho SS, Seruca R, Gärtner F, Yamaguchi Y, Gu J, Taniguchi N and Reis CA: Modulation of E-cadherin function and dysfunction by N-glycosylation. Cell Mol Life Sci. 68:1011–1020. 2011. View Article : Google Scholar

52 

Baschieri F, Confalonieri S, Bertalot G, Di Fiore PP, Dietmaier W, Leist M, Crespo P, Macara IG and Farhan H: Spatial control of Cdc42 signalling by a GM130-RasGRF complex regulates polarity and tumorigenesis. Nat Commun. 5:48392014. View Article : Google Scholar : PubMed/NCBI

53 

Taniguchi N and Kizuka Y: Glycans and cancer: Role of N-glycans in cancer biomarker, progression and metastasis, and therapeutics. Adv Cancer Res. 126:11–51. 2015. View Article : Google Scholar : PubMed/NCBI

54 

Rizzo R, Parashuraman S, D'Angelo G and Luini A: GOLPH3 and oncogenesis: What is the molecular link? Tissue Cell. 49:170–174. 2017. View Article : Google Scholar

55 

Farber-Katz SE, Dippold HC, Buschman MD, Peterman MC, Xing M, Noakes CJ, Tat J, Ng MM, Rahajeng J, Cowan DM, et al: DNA damage triggers Golgi dispersal via DNA-PK and GOLPH3. Cell. 156:413–427. 2014. View Article : Google Scholar : PubMed/NCBI

56 

Scott KL, Kabbarah O, Liang MC, Ivanova E, Anagnostou V, Wu J, Dhakal S, Wu M, Chen S, Feinberg T, et al: GOLPH3 modulates mTOR signalling and rapamycin sensitivity in cancer. Nature. 459:1085–1090. 2009. View Article : Google Scholar : PubMed/NCBI

57 

Xing M, Peterman MC, Davis RL, Oegema K, Shiau AK and Field SJ: GOLPH3 drives cell migration by promoting Golgi reorientation and directional trafficking to the leading edge. Mol Biol Cell. 27:3828–3840. 2016. View Article : Google Scholar : PubMed/NCBI

58 

Chang SH, Hong SH, Jiang HL, Minai-Tehrani A, Yu KN, Lee JH, Kim JE, Shin JY, Kang B, Park S, et al: GOLGA2/GM130, cis-Golgi matrix protein, is a novel target of anticancer gene therapy. Mol Ther. 20:2052–2063. 2012. View Article : Google Scholar : PubMed/NCBI

59 

Sehgal PB, Mukhopadhyay S, Patel K, Xu F, Almodóvar S, Tuder RM and Flores SC: Golgi dysfunction is a common feature in idiopathic human pulmonary hypertension and vascular lesions in SHIV-nef-infected macaques. Am J Physiol Lung Cell Mol Physiol. 297:L729–L737. 2009. View Article : Google Scholar : PubMed/NCBI

60 

Sehgal PB and Lee JE: Protein trafficking dysfunctions: Role in the pathogenesis of pulmonary arterial hypertension. Pulm Circ. 1:17–32. 2011. View Article : Google Scholar : PubMed/NCBI

61 

Muhammad E, Levitas A, Singh SR, Braiman A, Ofir R, Etzion S, Sheffield VC, Etzion Y, Carrier L and Parvari R: PLEKHM2 mutation leads to abnormal localization of lysosomes, impaired autophagy flux and associates with recessive dilated cardiomyopathy and left ventricular noncompaction. Hum Mol Genet. 24:7227–7240. 2015. View Article : Google Scholar : PubMed/NCBI

62 

Hatt PY: Cellular changes and damage in mechanically over-loaded hearts. Recent Adv Stud Cardiac Struct Metab. 6:325–333. 1975.

63 

Satoh H: Sino-atrial nodal cells of mammalian hearts: Ionic currents and gene expression of pacemaker ionic channels. J Smooth Muscle Res. 39:175–193. 2003. View Article : Google Scholar : PubMed/NCBI

64 

Jungk L, Franke H, Salameh A and Dhein S: Golgi fragmentation in human patients with chronic atrial fibrillation: A new aspect of remodeling. Thorac Cardiovasc Surg. 67:98–106. 2019. View Article : Google Scholar

65 

Prasad K and Singal PK: Ultrastructure of failing myocardium due to induced chronic mitral insufficiency in dogs. Br J Exp Pathol. 58:289–300. 1977.PubMed/NCBI

66 

Rambourg A, Clermont Y and Hermo L: Three-dimensional architecture of the golgi apparatus in sertoli cells of the rat. Am J Anat. 154:455–476. 1979. View Article : Google Scholar : PubMed/NCBI

67 

Xiang Y and Wang Y: New components of the Golgi matrix. Cell Tissue Res. 344:365–379. 2011. View Article : Google Scholar : PubMed/NCBI

68 

Kooy J, Toh BH, Pettitt JM, Erlich R and Gleeson PA: Human autoantibodies as reagents to conserved Golgi components. Characterization of a peripheral, 230-kDa compartment-specific Golgi protein. J Biol Chem. 267:20255–20263. 1992. View Article : Google Scholar : PubMed/NCBI

69 

Fritzler MJ, Hamel JC, Ochs RL and Chan EK: Molecular characterization of two human autoantigens: Unique cDNAs encoding 95- and 160-kD proteins of a putative family in the Golgi complex. J Exp Med. 178:49–62. 1993. View Article : Google Scholar : PubMed/NCBI

70 

Wong M and Munro S: Membrane trafficking. The specificity of vesicle traffic to the Golgi is encoded in the golgin coiled-coil proteins. Science. 346:12568982014. View Article : Google Scholar : PubMed/NCBI

71 

Barr FA, Puype M, Vandekerckhove J and Warren G: GRASP65, a protein involved in the stacking of Golgi cisternae. Cell. 91:253–262. 1997. View Article : Google Scholar : PubMed/NCBI

72 

Shorter J and Warren G: A role for the vesicle tethering protein, p115, in the post-mitotic stacking of reassembling Golgi cisternae in a cell-free system. J Cell Biol. 146:57–70. 1999. View Article : Google Scholar : PubMed/NCBI

73 

Vinke FP, Grieve AG and Rabouille C: The multiple facets of the Golgi reassembly stacking proteins. Biochem J. 433:423–433. 2011. View Article : Google Scholar : PubMed/NCBI

74 

Bergen DJM, Stevenson NL, Skinner REH, Stephens DJ and Hammond CL: The Golgi matrix protein giantin is required for normal cilia function in zebrafish. Biol Open. 6:1180–1189. 2017. View Article : Google Scholar : PubMed/NCBI

75 

Hennies HC, Kornak U, Zhang H, Egerer J, Zhang X, Seifert W, Kühnisch J, Budde B, Nätebus M, Brancati F, et al: Gerodermia osteodysplastica is caused by mutations in SCYL1BP1, a Rab-6 interacting golgin. Nat Genet. 40:1410–1412. 2008. View Article : Google Scholar : PubMed/NCBI

76 

Nakamura N, Rabouille C, Watson R, Nilsson T, Hui N, Slusarewicz P, Kreis TE and Warren G: Characterization of a cis-Golgi matrix protein, GM130. J Cell Biol. 131:1715–1726. 1995. View Article : Google Scholar : PubMed/NCBI

77 

Alvarez C, Garcia-Mata R, Hauri HP and Sztul E: The p115-inter- active proteins GM130 and giantin participate in endoplasmic reticulum-Golgi traffic. J Biol Chem. 276:2693–2700. 2001. View Article : Google Scholar

78 

Huang W, She L, Chang XY, Yang RR, Wang L, Ji HB, Jiao JW and Poo MM: Protein kinase LKB1 regulates polarized dendrite formation of adult hippocampal newborn neurons. Proc Natl Acad Sci USA. 111:469–474. 2014. View Article : Google Scholar

79 

Liu C, Mei M, Li Q, Roboti P, Pang Q, Ying Z, Gao F, Lowe M and Bao S: Loss of the golgin GM130 causes Golgi disruption, Purkinje neuron loss, and ataxia in mice. Proc Natl Acad Sci USA. 114:346–351. 2017. View Article : Google Scholar

80 

Matanis T, Akhmanova A, Wulf P, Del Nery E, Weide T, Stepanova T, Galjart N, Grosveld F, Goud B, De Zeeuw CI, et al: Bicaudal-D regulates COPI-independent Golgi-ER transport by recruiting the dynein-dynactin motor complex. Nat Cell Biol. 4:986–992. 2002. View Article : Google Scholar : PubMed/NCBI

81 

Hoogenraad CC, Akhmanova A, Howell SA, Dortland BR, De Zeeuw CI, Willemsen R, Visser P, Grosveld F and Galjart N: Mammalian Golgi-associated Bicaudal-D2 functions in the dynein-dynactin pathway by interacting with these complexes. EMBO J. 20:4041–4054. 2001. View Article : Google Scholar : PubMed/NCBI

82 

Jaarsma D, van den Berg R, Wulf PS, van Erp S, Keijzer N, Schlager MA, de Graaff E, De Zeeuw CI, Pasterkamp RJ, Akhmanova A and Hoogenraad CC: A role for bicaudal-D2 in radial cerebellar granule cell migration. Nat Commun. 5:34112014. View Article : Google Scholar : PubMed/NCBI

83 

Will L, Portegies S, van Schelt J, van Luyk M, Jaarsma D and Hoogenraad CC: Dynein activating adaptor BICD2 controls radial migration of upper-layer cortical neurons in vivo. Acta Neuropathol Commun. 7:1622019. View Article : Google Scholar : PubMed/NCBI

84 

Storbeck M, Horsberg Eriksen B, Unger A, Hölker I, Aukrust I, Martínez-Carrera LA, Linke WA, Ferbert A, Heller R, Vorgerd M, et al: Phenotypic extremes of BICD2-opathies: From lethal, congenital muscular atrophy with arthrogryposis to asymptomatic with subclinical features. Eur J Hum Genet. 25:1040–1048. 2017. View Article : Google Scholar : PubMed/NCBI

85 

Neveling K, Martinez-Carrera LA, Hölker I, Heister A, Verrips A, Hosseini-Barkooie SM, Gilissen C, Vermeer S, Pennings M, Meijer R, et al: Mutations in BICD2, which encodes a golgin and important motor adaptor, cause congenital autosomal-dominant spinal muscular atrophy. Am J Hum Genet. 92:946–954. 2013. View Article : Google Scholar : PubMed/NCBI

86 

Oates EC, Rossor AM, Hafezparast M, Gonzalez M, Speziani F, MacArthur DG, Lek M, Cottenie E, Scoto M, Foley AR, et al: Mutations in BICD2 cause dominant congenital spinal muscular atrophy and hereditary spastic paraplegia. Am J Hum Genet. 92:965–973. 2013. View Article : Google Scholar : PubMed/NCBI

87 

Sonnichsen B, Lowe M, Levine T, Jämsä E, Dirac-Svejstrup B and Warren G: A role for giantin in docking COPI vesicles to Golgi membranes. J Cell Biol. 140:1013–1021. 1998. View Article : Google Scholar : PubMed/NCBI

88 

Katayama K, Kuriki M, Kamiya T, Tochigi Y and Suzuki H: Giantin is required for coordinated production of aggrecan, link protein and type XI collagen during chondrogenesis. Biochem Biophys Res Commun. 499:459–465. 2018. View Article : Google Scholar : PubMed/NCBI

89 

Stevenson NL, Bergen DJM, Skinner REH, Kague E, Martin-Silverstone E, Robson Brown KA, Hammond CL and Stephens DJ: Giantin-knockout models reveal a feedback loop between Golgi function and glycosyltransferase expression. J Cell Sci. 130:4132–4143. 2017. View Article : Google Scholar : PubMed/NCBI

90 

Witkos TM, Chan WL, Joensuu M, Rhiel M, Pallister E, Thomas-Oates J, Mould AP, Mironov AA, Biot C, Guerardel Y, et al: GORAB scaffolds COPI at the trans-Golgi for efficient enzyme recycling and correct protein glycosylation. Nat Commun. 10:1272019. View Article : Google Scholar : PubMed/NCBI

91 

Sato K, Roboti P, Mironov AA and Lowe M: Coupling of vesicle tethering and Rab binding is required for in vivo functionality of the golgin GMAP-210. Mol Biol Cell. 26:537–553. 2015. View Article : Google Scholar :

92 

Smits P, Bolton AD, Funari V, Hong M, Boyden ED, Lu L, Manning DK, Dwyer ND, Moran JL, Prysak M, et al: Lethal skeletal dysplasia in mice and humans lacking the golgin GMAP-210. N Engl J Med. 362:206–216. 2010. View Article : Google Scholar : PubMed/NCBI

93 

Wehrle A, Witkos TM, Unger S, Schneider J, Follit JA, Hermann J, Welting T, Fano V, Hietala M, Vatanavicharn N, et al: Hypomorphic mutations of TRIP11 cause odontochondrodysplasia. JCI Insight. 4:e1247012019. View Article : Google Scholar :

94 

West DW: Energy-dependent calcium sequestration activity in a Golgi apparatus fraction derived from lactating rat mammary glands. Biochim Biophys Acta. 673:374–386. 1981. View Article : Google Scholar : PubMed/NCBI

95 

Shull GE, Miller ML and Prasad V: Secretory pathway stress responses as possible mechanisms of disease involving Golgi Ca2+ pump dysfunction. Biofactors. 37:150–158. 2011. View Article : Google Scholar : PubMed/NCBI

96 

Sudbrak R, Brown J, Dobson-Stone C, Carter S, Ramser J, White J, Healy E, Dissanayake M, Larrègue M, Perrussel M, et al: Hailey-Hailey disease is caused by mutations in ATP2C1 encoding a novel Ca(2+) pump. Hum Mol Genet. 9:1131–1140. 2000. View Article : Google Scholar : PubMed/NCBI

97 

Hu Z, Bonifas JM, Beech J, Bench G, Shigihara T, Ogawa H, Ikeda S, Mauro T and Epstein EH Jr: Mutations in ATP2C1, encoding a calcium pump, cause Hailey-Hailey disease. Nat Genet. 24:61–65. 2000. View Article : Google Scholar

98 

Okunade GW, Miller ML, Azhar M, Andringa A, Sanford LP, Doetschman T, Prasad V and Shull GE: Loss of the Atp2c1 secretory pathway Ca(2+)-ATPase (SPCA1) in mice causes Golgi stress, apoptosis, and midgestational death in homozygous embryos and squamous cell tumors in adult heterozygotes. J Biol Chem. 282:26517–26527. 2007. View Article : Google Scholar : PubMed/NCBI

99 

Lázaro-Diéguez F, Jiménez N, Barth H, Koster AJ, Renau-Piqueras J, Llopis JL, Burger KN and Egea G: Actin filaments are involved in the maintenance of Golgi cisternae morphology and intra-Golgi pH. Cell Motil Cytoskeleton. 63:778–791. 2006. View Article : Google Scholar : PubMed/NCBI

100 

Huang C and Chang A: pH-dependent cargo sorting from the Golgi. J Biol Chem. 286:10058–10065. 2011. View Article : Google Scholar : PubMed/NCBI

101 

Rivinoja A, Pujol FM, Hassinen A and Kellokumpu S: Golgi pH, its regulation and roles in human disease. Ann Med. 44:542–554. 2012. View Article : Google Scholar

102 

Drory O and Nelson N: The emerging structure of vacuolar ATPases. Physiology (Bethesda). 21. pp. 317–325. 2006

103 

Maeda Y, Ide T, Koike M, Uchiyama Y and Kinoshita T: GPHR is a novel anion channel critical for acidification and functions of the Golgi apparatus. Nat Cell Biol. 10:1135–1145. 2008. View Article : Google Scholar : PubMed/NCBI

104 

Banerjee S and Kane PM: Regulation of V-ATPase activity and organelle pH by phosphatidylinositol phosphate lipids. Front Cell Dev Biol. 8:5102020. View Article : Google Scholar : PubMed/NCBI

105 

Morava E, Guillard M, Lefeber DJ and Wevers RA: Autosomal recessive cutis laxa syndrome revisited. Eur J Hum Genet. 17:1099–1110. 2009. View Article : Google Scholar : PubMed/NCBI

106 

Van Damme T, Gardeitchik T, Mohamed M, Guerrero-Castillo S, Freisinger P, Guillemyn B, Kariminejad A, Dalloyaux D, van Kraaij S, Lefeber DJ, et al: Mutations in ATP6V1E1 or ATP6V1A cause autosomal-recessive cutis laxa. Am J Hum Genet. 100:216–227. 2017. View Article : Google Scholar : PubMed/NCBI

107 

Kariminejad A, Afroozan F, Bozorgmehr B, Ghanadan A, Akbaroghli S, Khorram Khorshid HR, Mojahedi F, Setoodeh A, Loh A, Tan YX, et al: Discriminative features in three autosomal recessive cutis laxa syndromes: Cutis laxa IIA, cutis laxa IIB, and geroderma osteoplastica. Int J Mol Sci. 18:6352017. View Article : Google Scholar :

108 

Kornak U, Reynders E, Dimopoulou A, van Reeuwijk J, Fischer B, Rajab A, Budde B, Nürnberg P, Foulquier F; ARCL Debré-type Study Group; et al: Impaired glycosylation and cutis laxa caused by mutations in the vesicular H+-ATPase subunit ATP6V0A2. Nat Genet. 40:32–34. 2008. View Article : Google Scholar

109 

Tümer Z: An overview and update of ATP7A mutations leading to Menkes disease and occipital horn syndrome. Hum Mutat. 34:417–429. 2013. View Article : Google Scholar : PubMed/NCBI

110 

Huster D, Hoppert M, Lutsenko S, Zinke J, Lehmann C, Mössner J, Berr F and Caca K: Defective cellular localization of mutant ATP7B in Wilson's disease patients and hepatoma cell lines. Gastroenterology. 124:335–345. 2003. View Article : Google Scholar : PubMed/NCBI

111 

Guan JL, Machamer CE and Rose JK: Glycosylation allows cell-surface transport of an anchored secretory protein. Cell. 42:489–496. 1985. View Article : Google Scholar : PubMed/NCBI

112 

Roth J: Protein N-glycosylation along the secretory pathway: Relationship to organelle topography and function, protein quality control, and cell interactions. Chem Rev. 102:285–303. 2002. View Article : Google Scholar : PubMed/NCBI

113 

Rosnoblet C, Peanne R, Legrand D and Foulquier F: Glycosylation disorders of membrane trafficking. Glycoconj J. 30:23–31. 2013. View Article : Google Scholar

114 

Grewal PK, McLaughlan JM, Moore CJ, Browning CA and Hewitt JE: Characterization of the LARGE family of putative glycosyltransferases associated with dystroglycanopathies. Glycobiology. 15:912–923. 2005. View Article : Google Scholar : PubMed/NCBI

115 

Kobayashi K, Nakahori Y, Miyake M, Matsumura K, Kondo-Iida E, Nomura Y, Segawa M, Yoshioka M, Saito K, Osawa M, et al: An ancient retrotransposal insertion causes Fukuyama-type congenital muscular dystrophy. Nature. 394:388–392. 1998. View Article : Google Scholar : PubMed/NCBI

116 

Brockington M, Blake DJ, Prandini P, Brown SC, Torelli S, Benson MA, Ponting CP, Estournet B, Romero NB, Mercuri E, et al: Mutations in the fukutin-related protein gene (FKRP) cause a form of congenital muscular dystrophy with secondary laminin alpha2 deficiency and abnormal glycosylation of alpha-dystroglycan. Am J Hum Genet. 69:1198–1209. 2001. View Article : Google Scholar : PubMed/NCBI

117 

Desplats PA, Denny CA, Kass KE, Gilmartin T, Head SR, Sutcliffe JG, Seyfried TN and Thomas EA: Glycolipid and ganglioside metabolism imbalances in Huntington's disease. Neurobiol Dis. 27:265–277. 2007. View Article : Google Scholar : PubMed/NCBI

118 

Cooper-Knock J, Moll T, Ramesh T, Castelli L, Beer A, Robins H, Fox I, Niedermoser I, Van Damme P, Moisse M, et al: Mutations in the glycosyltransferase domain of GLT8D1 are associated with familial amyotrophic lateral sclerosis. Cell Rep. 26:2298–2306.e5. 2019. View Article : Google Scholar : PubMed/NCBI

119 

Li W, Liu Z, Sun W, Yuan Y, Hu Y, Ni J, Jiao B, Fang L, Li J, Shen L, et al: Mutation analysis of GLT8D1 and ARPP21 genes in amyotrophic lateral sclerosis patients from mainland China. Neurobiol Aging. 85:156.e1–156.e4. 2020. View Article : Google Scholar

120 

Liu S and Storrie B: Are Rab proteins the link between Golgi organization and membrane trafficking? Cell Mol Life Sci. 69:4093–4106. 2012. View Article : Google Scholar : PubMed/NCBI

121 

Salian S, Cho TJ, Phadke SR, Gowrishankar K, Bhavani GS, Shukla A, Jagadeesh S, Kim OH, Nishimura G and Girisha KM: Additional three patients with Smith-McCort dysplasia due to novel RAB33B mutations. Am J Med Genet A. 173:588–595. 2017. View Article : Google Scholar : PubMed/NCBI

122 

Giannandrea M, Bianchi V, Mignogna ML, Sirri A, Carrabino S, D'Elia E, Vecellio M, Russo S, Cogliati F, Larizza L, et al: Mutations in the small GTPase gene RAB39B are responsible for X-linked mental retardation associated with autism, epilepsy, and macrocephaly. Am J Hum Genet. 86:185–195. 2010. View Article : Google Scholar : PubMed/NCBI

123 

Bock JB, Matern HT, Peden AA and Scheller RH: A genomic perspective on membrane compartment organization. Nature. 409:839–841. 2001. View Article : Google Scholar : PubMed/NCBI

124 

Corbett MA, Schwake M, Bahlo M, Dibbens LM, Lin M, Gandolfo LC, Vears DF, O'Sullivan JD, Robertson T, Bayly MA, et al: A mutation in the Golgi Qb-SNARE gene GOSR2 causes progressive myoclonus epilepsy with early ataxia. Am J Hum Genet. 88:657–663. 2011. View Article : Google Scholar : PubMed/NCBI

125 

Lowe SL, Peter F, Subramaniam VN, Wong SH and Hong W: A SNARE involved in protein transport through the Golgi apparatus. Nature. 389:881–884. 1997. View Article : Google Scholar : PubMed/NCBI

126 

Malsam J and Söllner TH: Organization of SNAREs within the Golgi stack. Cold Spring Harb Perspect Biol. 3:a0052492011. View Article : Google Scholar : PubMed/NCBI

127 

Gedeon AK, Colley A, Jamieson R, Thompson EM, Rogers J, Sillence D, Tiller GE, Mulley JC and Gécz J: Identification of the gene (SEDL) causing X-linked spondyloepiphyseal dysplasia tarda. Nat Genet. 22:400–404. 1999. View Article : Google Scholar : PubMed/NCBI

128 

Willett R, Ungar D and Lupashin V: The Golgi puppet master: COG complex at center stage of membrane trafficking interactions. Histochem Cell Biol. 140:271–283. 2013. View Article : Google Scholar : PubMed/NCBI

129 

Climer LK, Dobretsov M and Lupashin V: Defects in the COG complex and COG-related trafficking regulators affect neuronal Golgi function. Front Neurosci. 9:4052015. View Article : Google Scholar : PubMed/NCBI

130 

Miller VJ and Ungar D: Re'COG'nition at the Golgi. Traffic. 13:891–897. 2012. View Article : Google Scholar : PubMed/NCBI

131 

Egorov MV, Capestrano M, Vorontsova OA, Di Pentima A, Egorova AV, Mariggiò S, Ayala MI, Tetè S, Gorski JL, Luini A, et al: Faciogenital dysplasia protein (FGD1) regulates export of cargo proteins from the golgi complex via Cdc42 activation. Mol Biol Cell. 20:2413–2427. 2009. View Article : Google Scholar : PubMed/NCBI

132 

Roboti P, Sato K and Lowe M: The golgin GMAP-210 is required for efficient membrane trafficking in the early secretory pathway. J Cell Sci. 128:1595–1606. 2015. View Article : Google Scholar : PubMed/NCBI

133 

Zhang H, Winckler B and Cai Q: Introduction to the special issue on membrane trafficking in neurons. Dev Neurobiol. 78:167–169. 2018. View Article : Google Scholar : PubMed/NCBI

134 

Rosnoblet C, Legrand D, Demaegd D, Hacine-Gherbi H, de Bettignies G, Bammens R, Borrego C, Duvet S, Morsomme P, Matthijs G and Foulquier F: Impact of disease-causing mutations on TMEM165 subcellular localization, a recently identified protein involved in CDG-II. Hum Mol Genet. 22:2914–2928. 2013. View Article : Google Scholar : PubMed/NCBI

135 

Larson AA, Baker PR II, Milev MP, Press CA, Sokol RJ, Cox MO, Lekostaj JK, Stence AA, Bossler AD, Mueller JM, et al: TRAPPC11 and GOSR2 mutations associate with hypoglycosylation of α-dystroglycan and muscular dystrophy. Skelet Muscle. 8:172018. View Article : Google Scholar

136 

Davis EE, Savage JH, Willer JR, Jiang YH, Angrist M, Androutsopoulos A and Katsanis N: Whole exome sequencing and functional studies identify an intronic mutation in TRAPPC2 that causes SEDT. Clin Genet. 85:359–364. 2014. View Article : Google Scholar

137 

Riener MO: Diagnosis of tumours of the liver and the biliary tract: New tissue and serum markers. Pathologe. 32(Suppl 2): S304–S309. 2011.In German. View Article : Google Scholar

138 

Xu Z, Liu L, Pan X, Wei K, Wei M, Liu L, Yang H and Liu Q: Serum Golgi protein 73 (GP73) is a diagnostic and prognostic marker of chronic HBV liver disease. Medicine (Baltimore). 94:e6592015. View Article : Google Scholar

139 

Liu Y, Zou Z, Zhu B, Hu Z and Zeng P: CXCL10 decreases GP73 expression in hepatoma cells at the early stage of hepatitis C virus (HCV) infection. Int J Mol Sci. 14:24230–24241. 2013. View Article : Google Scholar : PubMed/NCBI

140 

Zheng KI, Liu WY, Pan XY, Ma HL, Zhu PW, Wu XX, Targher G, Byrne C, Wang XD, Chen YP, et al: Combined and sequential non-invasive approach to diagnosing non-alcoholic steatohepatitis in patients with non-alcoholic fatty liver disease and persistently normal alanine aminotransferase levels. BMJ Open Diabetes Res Care. 8:e0011742020. View Article : Google Scholar : PubMed/NCBI

141 

Hou SC, Xiao MB, Ni RZ, Ni WK, Jiang F, Li XY, Lu CH and Chen BY: Serum GP73 is complementary to AFP and GGT-II for the diagnosis of hepatocellular carcinoma. Oncol Lett. 6:1152–1158. 2013. View Article : Google Scholar : PubMed/NCBI

142 

Zhao J, Guo LY, Yang JM and Jia JW: Sublingual vein parameters, AFP, AFP-L3, and GP73 in patients with hepatocellular carcinoma. Genet Mol Res. 14:7062–7067. 2015. View Article : Google Scholar : PubMed/NCBI

143 

Liu X, Wan X, Li Z, Lin C, Zhan Y and Lu X: Golgi protein 73(GP73), a useful serum marker in liver diseases. Clin Chem Lab Med. 49:1311–1316. 2011. View Article : Google Scholar : PubMed/NCBI

144 

Morota K, Nakagawa M, Sekiya R, Hemken PM, Sokoll LJ, Elliott D, Chan DW and Dowell BL: A comparative evaluation of Golgi protein-73, fucosylated hemopexin, α-fetoprotein, and PIVKA-II in the serum of patients with chronic hepatitis, cirrhosis, and hepatocellular carcinoma. Clin Chem Lab Med. 49:711–718. 2011. View Article : Google Scholar : PubMed/NCBI

145 

Gu Y, Chen W, Zhao Y, Chen L and Peng T: Quantitative analysis of elevated serum Golgi protein-73 expression in patients with liver diseases. Ann Clin Biochem. 46:38–43. 2009. View Article : Google Scholar

146 

Tian L, Wang Y, Xu D, Gui J, Jia X, Tong H, Wen X, Dong Z and Tian Y: Serological AFP/Golgi protein 73 could be a new diagnostic parameter of hepatic diseases. Int J Cancer. 129:1923–1931. 2011. View Article : Google Scholar

147 

Ye JZ, Yan SM, Yuan CL, Wu HN, Zhang JY, Liu ZH, Li YQ, Luo XL, Lin Y and Liang R: GP73 level determines chemo-therapeutic resistance in human hepatocellular carcinoma cells. J Cancer. 9:415–423. 2018. View Article : Google Scholar :

148 

Lebredonchel E, Houdou M, Potelle S, de Bettignies G, Schulz C, Krzewinski Recchi MA, Lupashin V, Legrand D, Klein A and Foulquier F: Dissection of TMEM165 function in Golgi glycosylation and its Mn2+ sensitivity. Biochimie. 165:123–130. 2019. View Article : Google Scholar : PubMed/NCBI

149 

Lee JS, Kim MY, Park ER, Shen YN, Jeon JY, Cho EH, Park SH, Han CJ, Choi DW, Jang JJ, et al: TMEM165, a Golgi transmembrane protein, is a novel marker for hepatocellular carcinoma and its depletion impairs invasion activity. Oncol Rep. 40:1297–1306. 2018.PubMed/NCBI

150 

Lee SH, Yoo HJ, Rim DE, Cui Y, Lee A, Jung ES, Oh ST, Kim JG, Kwon OJ, Kim SY and Jeong SW: Nuclear expression of GS28 protein: A novel biomarker that predicts prognosis in colorectal cancers. Int J Med Sci. 14:515–522. 2017. View Article : Google Scholar : PubMed/NCBI

151 

Cho U, Kim HM, Park HS, Kwon OJ, Lee A and Jeong SW: Nuclear expression of GS28 protein: A novel biomarker that predicts worse prognosis in cervical cancers. PLoS One. 11:e01626232016. View Article : Google Scholar : PubMed/NCBI

152 

Rodriguez JL, Gelpi C, Thomson TM, Real FJ and Fernández J: Anti-golgi complex autoantibodies in a patient with Sjögren syndrome and lymphoma. Clin Exp Immunol. 49:579–586. 1982.

153 

Fritzler MJ, Etherington J, Sokoluk C, Kinsella TD and Valencia DW: Antibodies from patients with autoimmune disease react with a cytoplasmic antigen in the Golgi apparatus. J Immunol. 132:2904–2908. 1984.PubMed/NCBI

154 

Hong HS, Morshed SA, Tanaka S, Fujiwara T, Ikehara Y and Nishioka M: Anti-Golgi antibody in rheumatoid arthritis patients recognizes a novel antigen of 79 kDa (doublet) by western blot. Scand J Immunol. 36:785–792. 1992. View Article : Google Scholar : PubMed/NCBI

155 

Gentric A, Blaschek M, Julien C, Jouquan J, Pennec Y, Berthelot JM, Mottier D, Casburn-Budd R and Youinou P: Nonorgan-specific autoantibodies in individuals infected with type 1 human immunodeficiency virus. Clin Immunol Immunopathol. 59:487–494. 1991. View Article : Google Scholar : PubMed/NCBI

156 

Griffith KJ, Chan EK, Lung CC, Hamel JC, Guo X, Miyachi K and Fritzler MJ: Molecular cloning of a novel 97-kd Golgi complex autoantigen associated with Sjögren's syndrome. Arthritis Rheum. 40:1693–1702. 1997. View Article : Google Scholar : PubMed/NCBI

157 

Gaspar ML, Marcos MA, Gutierrez C, Martin MJ, Bonifacino JS and Sandoval IV: Presence of an autoantibody against a Golgi cisternal membrane protein in the serum and cerebrospinal fluid from a patient with idiopathic late onset cerebellar ataxia. J Neuroimmunol. 17:287–299. 1988. View Article : Google Scholar : PubMed/NCBI

158 

Huidbüchel E, Blaschek M, Seigneurin JM, Lamour A, Berthelot JM and Youinou P: Anti-organelle and anti-cytoskeletal autoantibodies in the serum of Epstein-Barr virus-infected patients. Ann Med Interne (Paris). 142:343–346. 1991.

159 

Paraná R, Schinoni MI, de Freitas LA, Codes L, Cruz M, Andrade Z and Trepo C: Anti-Golgi complex antibodies during pegylated-interferon therapy for hepatitis C. Liver Int. 26:1148–1154. 2006. View Article : Google Scholar : PubMed/NCBI

160 

Mozo L, Simó A, Suarez A, Rodrigo L and Gutiérrez C: Autoantibodies to Golgi proteins in hepatocellular carcinoma: Case report and literature review. Eur J Gastroenterol Hepatol. 14:771–774. 2002. View Article : Google Scholar : PubMed/NCBI

161 

Fritzler MJ, Hudson M, Choi MY, Mahler M, Wang M, Bentow C, Milo J and Baron M; Canadian Scleroderma Research Group: Bicaudal D2 is a novel autoantibody target in systemic sclerosis that shares a key epitope with CENP-A but has a distinct clinical phenotype. Autoimmun Rev. 17:267–275. 2018. View Article : Google Scholar : PubMed/NCBI

162 

Xue F, Wen Y, Wei P, Gao Y, Zhou Z, Xiao S and Yi T: A smart drug: A pH-responsive photothermal ablation agent for Golgi apparatus activated cancer therapy. Chem Commun (Camb). 53:6424–6427. 2017. View Article : Google Scholar

163 

Li H, Zhang P, Luo J, Hu D, Huang Y, Zhang ZR, Fu Y and Gong T: Chondroitin sulfate-linked prodrug nanoparticles target the Golgi apparatus for cancer metastasis treatment. ACS Nano. 13:9386–9396. 2019. View Article : Google Scholar : PubMed/NCBI

164 

Zappa F, Failli M and De Matteis MA: The Golgi complex in disease and therapy. Curr Opin Cell Biol. 50:102–116. 2018. View Article : Google Scholar : PubMed/NCBI

165 

Borck G, Mollà-Herman A, Boddaert N, Encha-Razavi F, Philippe A, Robel L, Desguerre I, Brunelle F, Benmerah A, Munnich A and Colleaux L: Clinical, cellular, and neuropathological consequences of AP1S2 mutations: Further delineation of a recognizable X-linked mental retardation syndrome. Hum Mutat. 29:966–974. 2008. View Article : Google Scholar : PubMed/NCBI

166 

Ammann S, Schulz A, Krägeloh-Mann I, Dieckmann NM, Niethammer K, Fuchs S, Eckl KM, Plank R, Werner R, Altmüller J, et al: Mutations in AP3D1 associated with immunodeficiency and seizures define a new type of Hermansky-Pudlak syndrome. Blood. 127:997–1006. 2016. View Article : Google Scholar : PubMed/NCBI

167 

Lu J, Tiao G, Folkerth R, Hecht J, Walsh C and Sheen V: Overlapping expression of ARFGEF2 and Filamin A in the neuroependymal lining of the lateral ventricles: Insights into the cause of periventricular heterotopia. J Comp Neurol. 494:476–484. 2006. View Article : Google Scholar

168 

Vogt G, El Choubassi N, Herczegfalvi Á, Kölbel H, Lekaj A, Schara U, Holtgrewe M, Krause S, Horvath R, Schuelke M, et al: Expanding the clinical and molecular spectrum of ATP6V1A related metabolic cutis laxa. J Inherit Metab Dis. Dec 15–2020.Online ahead of print. PubMed/NCBI

169 

Pflieger LT, Dansithong W, Paul S, Scoles DR, Figueroa KP, Meera P, Otis TS, Facelli JC and Pulst SM: Gene co-expression network analysis for identifying modules and functionally enriched pathways in SCA2. Hum Mol Genet. 26:3069–3080. 2017. View Article : Google Scholar : PubMed/NCBI

170 

Shestakova A, Zolov S and Lupashin V: COG complex-mediated recycling of Golgi glycosyltransferases is essential for normal protein glycosylation. Traffic. 7:191–204. 2006. View Article : Google Scholar : PubMed/NCBI

171 

Jensson BO, Hansdottir S, Arnadottir GA, Sulem G, Kristjansson RP, Oddsson A, Benonisdottir S, Jonsson H, Helgason A, Saemundsdottir J, et al: COPA syndrome in an Icelandic family caused by a recurrent missense mutation in COPA. BMC Med Genet. 18:1292017. View Article : Google Scholar : PubMed/NCBI

172 

Han C, Alkhater R, Froukh T, Minassian AG, Galati M, Liu RH, Fotouhi M, Sommerfeld J, Alfrook AJ, Marshall C, et al: Epileptic encephalopathy caused by mutations in the guanine nucleotide exchange factor DENND5A. Am J Hum Genet. 99:1359–1367. 2016. View Article : Google Scholar : PubMed/NCBI

173 

Dupuis N, Fafouri A, Bayot A, Kumar M, Lecharpentier T, Ball G, Edwards D, Bernard V, Dournaud P, Drunat S, et al: Dymeclin deficiency causes postnatal microcephaly, hypomyelination and reticulum-to-Golgi trafficking defects in mice and humans. Hum Mol Genet. 24:2771–2783. 2015. View Article : Google Scholar : PubMed/NCBI

174 

Utine GE, Taşkıran EZ, Koşukcu C, Karaosmanoğlu B, Güleray N, Doğan ÖA, Kiper PÖ, Boduroğlu K and Alikaşifoğlu M: HERC1 mutations in idiopathic intellectual disability. Eur J Med Genet. 60:279–283. 2017. View Article : Google Scholar : PubMed/NCBI

175 

Xing G, Yao J, Wu B, Liu T, Wei Q, Liu C, Lu Y, Chen Z, Zheng H, Yang X and Cao X: Identification of OSBPL2 as a novel candidate gene for progressive nonsyndromic hearing loss by whole-exome sequencing. Genet Med. 17:210–218. 2015. View Article : Google Scholar

176 

Dupuis N, Lebon S, Kumar M, Drunat S, Graul-Neumann LM, Gressens P and El Ghouzzi V: A novel RAB33B mutation in Smith-McCort dysplasia. Hum Mutat. 34:283–286. 2013. View Article : Google Scholar

177 

Kondo Y, Fu J, Wang H, Hoover C, McDaniel JM, Steet R, Patra D, Song J, Pollard L, Cathey S, et al: Site-1 protease deficiency causes human skeletal dysplasia due to defective inter-organelle protein trafficking. JCI Insight. 3:e1215962018. View Article : Google Scholar :

178 

Carvalho DR, Speck-Martins CE, Brum JM, Ferreira CR and Sobreira NLM: Spondyloepimetaphyseal dysplasia with elevated plasma lysosomal enzymes caused by homozygous variant in MBTPS1. Am J Med Genet A. 182:1796–1800. 2020. View Article : Google Scholar : PubMed/NCBI

179 

Ng BG, Asteggiano CG, Kircher M, Buckingham KJ, Raymond K, Nickerson DA, Shendure J, Bamshad MJ; University of Washington Center for Mendelian Genomics; Ensslen M and Freeze HH: Encephalopathy caused by novel mutations in the CMP-sialic acid transporter, SLC35A1. Am J Med Genet A. 173:2906–2911. 2017. View Article : Google Scholar : PubMed/NCBI

180 

Dörre K, Olczak M, Wada Y, Sosicka P, Grüneberg M, Reunert J, Kurlemann G, Fiedler B, Biskup S, Hörtnagel K, et al: A new case of UDP-galactose transporter deficiency (SLC35A2-CDG): Molecular basis, clinical phenotype, and therapeutic approach. J Inherit Metab Dis. 38:931–940. 2015. View Article : Google Scholar : PubMed/NCBI

181 

Wei J, Zhang YY, Luo J, Wang JQ, Zhou YX, Miao HH, Shi XJ, Qu YX, Xu J, Li BL and Song BL: The GARP complex is involved in intracellular cholesterol transport via targeting NPC2 to lysosomes. Cell Rep. 19:2823–2835. 2017. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2021
Volume 47 Issue 4

Print ISSN: 1107-3756
Online ISSN:1791-244X

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Liu J, Huang Y, Li T, Jiang Z, Zeng L and Hu Z: The role of the Golgi apparatus in disease (Review). Int J Mol Med 47: 38, 2021
APA
Liu, J., Huang, Y., Li, T., Jiang, Z., Zeng, L., & Hu, Z. (2021). The role of the Golgi apparatus in disease (Review). International Journal of Molecular Medicine, 47, 38. https://doi.org/10.3892/ijmm.2021.4871
MLA
Liu, J., Huang, Y., Li, T., Jiang, Z., Zeng, L., Hu, Z."The role of the Golgi apparatus in disease (Review)". International Journal of Molecular Medicine 47.4 (2021): 38.
Chicago
Liu, J., Huang, Y., Li, T., Jiang, Z., Zeng, L., Hu, Z."The role of the Golgi apparatus in disease (Review)". International Journal of Molecular Medicine 47, no. 4 (2021): 38. https://doi.org/10.3892/ijmm.2021.4871