Next Article in Journal
Changing the History of Prostate Cancer with New Targeted Therapies
Next Article in Special Issue
The Multifaceted Roles of Zinc in Neuronal Mitochondrial Dysfunction
Previous Article in Journal
Preoperative Localization for Primary Hyperparathyroidism: A Clinical Review
Previous Article in Special Issue
Dietary Iron Intake in Excess of Requirements Impairs Intestinal Copper Absorption in Sprague Dawley Rat Dams, Causing Copper Deficiency in Suckling Pups
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

ATP7A-Regulated Enzyme Metalation and Trafficking in the Menkes Disease Puzzle

by
Nina Horn
1,*,† and
Pernilla Wittung-Stafshede
2
1
John F. Kennedy Institute, 2600 Glostrup, Denmark
2
Department of Biology and Biological Engineering, Chalmers University of Technology, 41296 Gothenburg, Sweden
*
Author to whom correspondence should be addressed.
Retired.
Biomedicines 2021, 9(4), 391; https://doi.org/10.3390/biomedicines9040391
Submission received: 13 March 2021 / Revised: 1 April 2021 / Accepted: 2 April 2021 / Published: 6 April 2021
(This article belongs to the Special Issue Zinc and Copper in Human Health and Disease)

Abstract

:
Copper is vital for numerous cellular functions affecting all tissues and organ systems in the body. The copper pump, ATP7A is critical for whole-body, cellular, and subcellular copper homeostasis, and dysfunction due to genetic defects results in Menkes disease. ATP7A dysfunction leads to copper deficiency in nervous tissue, liver, and blood but accumulation in other tissues. Site-specific cellular deficiencies of copper lead to loss of function of copper-dependent enzymes in all tissues, and the range of Menkes disease pathologies observed can now be explained in full by lack of specific copper enzymes. New pathways involving copper activated lysosomal and steroid sulfatases link patient symptoms usually related to other inborn errors of metabolism to Menkes disease. Additionally, new roles for lysyl oxidase in activation of molecules necessary for the innate immune system, and novel adapter molecules that play roles in ERGIC trafficking of brain receptors and other proteins, are emerging. We here summarize the current knowledge of the roles of copper enzyme function in Menkes disease, with a focus on ATP7A-mediated enzyme metalation in the secretory pathway. By establishing mechanistic relationships between copper-dependent cellular processes and Menkes disease symptoms in patients will not only increase understanding of copper biology but will also allow for the identification of an expanding range of copper-dependent enzymes and pathways. This will raise awareness of rare patient symptoms, and thus aid in early diagnosis of Menkes disease patients.

1. ATP7A-Related Copper Disorders

ATP7A-related X-linked genetic disturbances exhibit dysfunction of multiple copper-dependent processes resulting in a broad spectrum of disease phenotypes. Three clinical groups are described: Menkes disease (MNK), occipital horn syndrome (OHS), and X-linked distal spinal muscular atrophy 3 (SMAX3) but overlapping intermediate forms (Table 1) confuse grouping [1,2]. MNK is characterized by neurodegeneration, fair skin, kinky hair, connective tissue abnormalities, and short life span. OHS presents with connective tissue symptoms, develops pathognomonic occipital bony exostosis (horns), and has reduced life expectancy. SMAX3 comprises a yet limited group of adult-onset progressive motor neuron disease, minimal copper disturbance, normal fertility, and long lifespan. Grouping into three phenotypes is arbitrary, and the spectrum is better described as a clinical continuum from severe disease with many affected enzyme systems to very mild affection with few enzyme systems involved. The best nosology should refer to ATP7A-related disturbances as the main pointer [2]. To best explain patients’ symptoms, we have here focused on the severest form, i.e., MNK.
The basic defect in MNK is deficient copper transfer to secretory pathway, the intracellular sorting station for proteins, which affects metalation and trafficking of copper-dependent enzymes and diminishes copper extrusion from cells. Insufficient functional ATP7A results in abnormal body copper distribution with high values in tissues, but severe lack in blood, brain, and liver [1]. Current understanding of copper dependent processes does not account for all neurological symptoms [3], and a broad update is needed. Substrates are included with each enzyme to better define key symptoms in MNK.
Clinical features found in MNK patients can be explained by deficiencies of copper-dependent enzymes, amine oxidases (biodegradation of histamine and polyamines), lysyl oxidases (cross-linking of elastin, collagen, and collectin), cytochrome c oxidase (energy formation), peptidyl α-amidating enzyme (activation of neurohormones and neuropeptides), dopamine β-hydroxylase (catecholamine production), tyrosinase (pigment formation and free radical defense), superoxide dismutase (free radical detoxification), and ferroxidase (iron mobilization and free radical defense). In addition, newly discovered copper-activated enzymes, including several lysosomal and steroid sulfatases, provide further insight into MNK pathophysiology [4] and will be discussed herein.

2. Copper Enzymes

Copper serves as redox cofactor in a wide range of reactions including electron transfer, oxidation, reduction, and disproportionation [4] (Table 2). Copper is regulated by redox shifts at several cellular and subcellular membranes and is further implicated in redox shifts of iron. In addition, copper acts as allosteric regulator at binding sites spatially apart from catalytic centers (Table 2). In enzymes, copper acts as integral electron modulator, but roles in formation and coupling of cofactors are important for understanding MNK clinical spectrum [4] and copper chaperone activation of enzymes without catalytic copper sites is emerging [5]. Copper enzymes contain active centers shuttling electrons from one molecule to another, and copper changes reversibly between oxidation states during catalytic cycles. Nature uses a variety of copper centers to facilitate electron transfer, and enzymes are grouped accordingly. Often catalytic metal sites are made up of metal coordinating residues located far apart in primary structure, but in the folded structure, they create a cluster of closely spaced amino acids forming metal-binding sites. The protein’s tertiary structure is often stabilized by coordination of the metal [6,7].
Copper enzymes are widely distributed both intracellularly and extracellularly. Intracellular copper enzymes are located in subcellular compartments and organelles such as cytoplasm, secretory pathway, peroxisomes, nucleus, and mitochondria, and metalation occurs at different sites. It is an important question to ask when and where the polypeptide meets the metal for copper-dependent enzymes. Newly synthesized apoenzymes are directed to the secretory pathway, and during passage from endoplasmic reticulum (ER) to trans-Golgi Network (TGN), maturation, assembly, glycosylation, and metalation occur. Copper transfer uses different mechanisms involving specific copper chaperones and combine metal coordination chemistry with protein-protein interactions for donor-acceptor docking. Specific copper chaperones for all copper-dependent enzymes have not been identified, and additional ones are likely to be discovered in the future. Cellular copper redox states and concentrations are strictly controlled, and free copper ions are kept at low, non-toxic levels. Within secretory pathway pH is gradually shifted towards an acidic environment and from oxidizing to reducing milieu changing the strength of copper binding [8,9]. Centers with tight metal coordination are preserved during secretory passage, while centers with low avidity can lose copper, and chaperones may be needed to protect their lability. Several copper enzymes possess labile copper sites, such as SOD1, SOD3, DBH, PAM, and TYR, while CP and HEPH have more stable copper sites.
This review centers on function and biogenesis of copper-dependent enzymes, which are activated/matured by copper loading at various sites in cells, which relates to protein trafficking. Cellular copper homeostasis is extensively covered in other reviews, e.g., [10,11,12], and will not be discussed here unless pertinent for understanding activation of copper enzymes.

3. Copper-Dependent ATPases

Copper-transporting ATPases, α (ATP7A) and β (ATP7B), are homologous P-type ATPases utilizing energy for pumping copper across a membrane. They are ion gated channels crucial for cellular and whole-body copper homeostasis. They have similar but distinct functions, and supplement and complement each other to fine tune equilibrium by transporting copper in different tissues and by coordinating activity in specific cells [11,13]. Both enzymes pump copper from cytoplasm into compartments with higher copper concentration [14]. ATP7A moves copper out of cytosol and across the basolateral membrane in extra-hepatic tissues [15], while ATP7B moves copper out of cytosol and across the apical membrane in liver, brain, and kidney [11,16]. ATP7A controls transport across the gut mucosa and the blood–brain barrier (BBB).
ATP7A and ATP7B are multi-domain enzymes that undergo profound changes during pumping [17]. They share highly conserved domain structure and basic mechanism with other P-type ATPases. Eight membrane-spanning helices constitute a pore-forming transmembrane domain for copper translocation. The channel is linked to three cytoplasmic ATP hydrolytic domains plus six metal binding domains (MBD) with copper-specific motifs (GMXCXXC) [18]. MBD’s initiate pumping through ATOX1 copper activation [11]. Each MBD possesses a compact fold linked by a flexible loop, enabling independent and cooperative action [19]. During pumping conformation undergoes a flip-flop movement with sequential changes allowing unidirectional transfer of copper from the entry site, through the channel by two embedded sites, and release from an exit site. A kinked transmembrane segment at the cytosolic interface forms an electronegative platform for electrostatic ATOX1 docking, initiating opening of the entry gate [11,20].
Pump domain interactions depend on conformation and position during the catalytic cycle, and energy for pumping stems from ATP-dependent transient phosphorylation of the cytoplasmic part [13]. Disruption of the cycle at any point reduces copper transfer.
Regulatory mechanisms are slowly unraveled. ATP7A/B pumping activity is via MBD’s controlled by copper. Docking of ATOX1 on the kinked platform, filling and packing of MBD’s serve as metal sensor besides allosteric regulation. Exact molecular mechanisms that modulate ATP7A/B activity still remain unclear [21].
ATP7A/B trafficking is copper regulated. At basic homeostatic levels copper is pumped into the secretory pathway, but at high levels the pumps relocate to excrete surplus. At low copper, a high free GSH pool secure glutathionylation of MBD’s and retention, while high copper results in low glutathionylation and trafficking [22,23]. ATP7A/B contain a histidine and methionine rich lumenal loop located between TM1 and TM2 that may function as ER retention signal [20,24,25,26]. Ca-pumps possess similar regulatory motifs at corresponding locations to secure Ca-guided ER retention [27]. The ATP7A loop motif may act as copper donor in metalation of certain enzymes [28]. ATP7A and ATP7B show distinct copper transport kinetics, where ATP7A is faster than ATP7B [11,29], but underlying reasons for differences are not clear [21].

3.1. Copper-Transporting ATPase 1 (ATP7A)

ATP7A regulates tissue copper levels and is expressed in most tissues except postnatal liver. At basal levels ATP7A transports copper into lumen of secretory pathway to load secreted and vesicular copper enzymes. At standard tissue culture conditions ATP7A reside in TGN and when exposed to excess copper, the pump relocates to the plasma membrane to export copper [30]. Some enzymes require metalation in ER, and removal of TGN signal by skipping of the alternatively spliced exon 10 retains the protein in ER [20,24] and may be of functional significance. ATP7A is rate limiting in gut uptake and import to the brain.
ATP7A interacts with a range of adaptor molecules, some affect nerve development [31]. ATP7A contains several N-glycosylation sites [32] and need ERGIC trafficking by a carbohydrate-recognition domain (CRD) and an adaptor complex like LMAN1 [33].

3.2. Copper-Transporting ATPase 2 (ATP7B)

ATP7B regulates whole-body copper homeostasis by excreting surplus into bile [34]. ATP7B is expressed in numerous tissues, and plays a role in copper regulation in liver, brain, placenta, and kidneys [4]. ATP7B supplies copper in liver to ceruloplasmin, and clotting factors V and VIII. Gene defects cause Wilson disease (WND) with copper accumulation in liver and brain.
ATP7B is not N-glycosylated [32] and not sugar sorted to the apical canalicular membrane but uses a subset of secretory lysosomes. Trafficking is directed by a motif of aromatic amino acids between MBD4 and MBD5 with loose copper binding [15,35] requiring an acidic milieu to secure high free copper for activation. Consistently ATP7B uses an acidic lysosomal pool for excretion. Inactivation of the trafficking signal directs ATP7B to the sinusoidal membrane [35] to mobilize hepatic stores into circulation [36]. In the brain, ATP7B likely traffics by other mechanisms, though not described. ATP7B is also regulated by alternative splicing of the loop motif [13,37].

4. Redox Shifting Enzymes

Intracellularly cupro ions, Cu(I) predominate, while higher extracellular oxidation potential results in cupric ions, Cu(II) [11,38]. A redox shift is needed at copper uptake and export, and organelle membranes likely also require copper redox shifts for transfer. This mimics conclusions for iron transport [39] and at sites the two metals share redox enzymes [40,41]. Copper and iron are reduced at plasma membranes by a heme reductase [42]. Before release iron uses a multicopper oxidase also having copper oxidizing capacity [43]. Iron has higher reduction potential than copper, while copper is superior in oxidative reactions [44,45]. Iron prosthetic groups are involved in a broad range of biological processes. Iron utility depends on careful control of redox state, and specific redox enzymes are found widespread also at subcellular levels. Iron–copper interactions have emerged as crucial, and copper is critical for normal handling of iron [46].
Mitochondria have significant iron and copper stores, securing biogenesis of two iron prosthetic groups, heme and iron-sulfur (Fe-S). Heme is tightly interconnected with copper metabolism and dependent on Fe-S availability [47]. Fe-S clusters are found at several subcellular sites including mitochondrial respiratory chain. Fe-S biogenesis is complex involving numerous steps, some occur in mitochondrial matrix, others in cytoplasm. Fe-S biosynthesis is interconnected with heme biosynthesis [48]. Iron trafficking is not well understood [49], and redox changes critical in organelle metal homeostasis are less known. Copper deficiency leads to low heme-iron which in turn gives insufficiency of enzymes needed for mitochondrial iron membrane translocation [50]. Deficient heme affects copper through dysfunction of membrane uptake, conferring a gatekeeping role for copper in translocation of both iron and copper.

4.1. Heme Copper Reductases

Six-transmembrane epithelial antigen of prostate (STEAP) comprises a family of metalloreductases with ability to reduce iron and copper [51]. STEAP4 shows physiological Km values for both metals [52]. Reducing sites use heme and cofactors like NAD, NAD(P)H, and ascorbate as electron donor [53,54]. The STEAP family is widely expressed [51], but tissue-specific expressions suggest distinct roles [53]. STEAP proteins locate at plasma membrane for copper and iron uptake but are also implicated in trafficking by modulating redox states in endocytotic and secretory pathway, and in mitochondria. STEAP1 acts at tight junctions, gap junctions, and cellular adhesion, and is hormone regulated [53]. STEAP2 regulates iron and copper absorption in gastrointestinal tract [53] and flux across BBB [55]. STEAP2 is expressed in most tissues primarily at plasma membrane and Golgi complex, possibly regulating metal availability in secretory pathway [53]. STEAP3 is an endosomal reductase required for efficient iron uptake into erythroid precursor cells. STEAP3 is highly expressed in liver, placenta, and bone marrow [53], and is located at plasma membrane, near nucleus, and in vesicular tubular structures [53]. STEAP4 is ubiquitously expressed at plasma membrane, ER, and TGN, suggesting a trafficking role [53,56]. STEAP4 also localizes at early endosomes and mitochondria [57,58], and splice variants localize to nucleus [59].
Cytochrome B reductase 1 (CYBRD1) is a di-heme reductase with iron and copper reducing capacity located at intestinal brush border to reduce iron and copper for mucosal uptake [60]. It may act also as a reductase in airway epithelium. CYBRD1 reduces iron and copper, uses ascorbate, and possesses 6TM, hence naturally grouped among STEAP.

4.2. Multicopper Oxidases

Multicopper oxidases (blue copper oxidases) contain six copper in two complex sites with different redox properties, oxidizing sequentially without formation of free radicals. Cupro and ferro ions are strong pro-oxidants, and multicopper oxidases scavenge radicals by preventing Fenton chemistry [11,43].

4.2.1. Ceruloplasmin (CP)

Ceruloplasmin (CP) is a major redox buffer in blood converting highly toxic ferro ions to less toxic ferri ions. CP mobilizes iron from stores in liver and other tissues, and copper by securing oxidation state [11,61], not by moving copper from one binding site to another; CP is not part of the exchangeable copper pool. Fet3p, a yeast homologue, exhibits cuproxidase activity at same site having ferroxidase activity [61]. CP is synthesized in liver, but a glycosylphosphatidylinositol-linked form (GPI-CP) is found in astrocytes and choroid plexus [62,63]. GPI is attached in ER and serves as O-glycosylation signal for trafficking to the apical membrane [62,64]. CP is ascorbate dependent and has two deeply embedded catalytic centers [6,65,66], copper loaded in liver by ATP7B, and released to circulation from the sinusoidal membrane. Apo-CP passes to Golgi though exact metalation site is not identified [65,67]. The apoform is secreted into blood, but rapidly degraded [68]. It lacks ferroxidase activity [65], and cannot be copper activated later [66], underlining that CP is not a transport form for tissue copper exchange [11]. Aceruloplasminemia leads to iron accumulation in brain and Parkinson-like ataxia and progressive dementia [69]. In ATP7A-related disturbances, CP metalation is not affected, but low hepatic copper results in low plasma holoenzyme activity. Ferroxidase activity is also low in other tissues, especially brain GPI-CP. Low plasma CP is well-documented in MNK, and some patients show moderate hypochromic anemia, and iron accumulates in kidney and nerve tissue [70]. Poor glycosylation may contribute to poor hepatic secretion, and MNK plasma CP is lower (half) than nutritional copper deficiency and WND [71]. Low ceruloplasmin is a well-established marker for MNK, OHS, and intermediate forms, though less depressed in OHS; but it is not applicable for diagnosis of SMAX3 [2].

4.2.2. Hephaestin (HEPH)

Hephaestin (HEPH) is a membrane-anchored homologue oxidizing iron before release with main role in gut and BBB [64]. Copper oxidizing capacity is likely [11]. The iron metabolome contains large functional redundancy and potentially CP and HEPH can substitute for each other [72], and CP may in absence of HEPH promote iron efflux from enterocytes [73].
HEPH and CP both appear critical for CNS-iron homeostasis [74], and HEPH is abundantly expressed in neurons [75]. CP and HEPH are co-expressed in retina [76] and combined loss lead to age-related macular degeneration (AMD) [72]. Double-knockout Heph and Cp mice, but neither alone, lead to kidney iron deposition and toxicity [77], and iron accumulation in liver, brain, pancreas, and adipocytes [74,78,79], with significant deficiencies in serum and neurons [80]. Knockout mice exhibit neurodegeneration and retinal degeneration similar to aceruloplasminemic patients. HEPH sites have similar architecture as CP but are loaded by ATP7A [43]. The apoform is rapidly degraded, and by analogy cannot be loaded after biosynthesis [81]. HEPH is regulated by copper and iron; iron induces translocation from intracellular sites to basolateral membrane [82]. HEPH is ubiquitously expressed, most strongly in small intestine followed by kidney [43]. HEPH is N-glycosylated [81], sorted and anchored to vesicles at basolateral membrane [64,81,83].

4.2.3. Hephaestin-Like Protein 1 (HEPHL1)

Hephaestin-like protein 1 (HEPHL1) is a membrane-bound homologue (zyklopen) with similar oxidation functions [84]. Immunostaining shows expression in brain, kidney, testes, and retina, but not liver and intestine [84]. Further expressed in placenta, reproductive tract, and mammary glands and suggested to act in placental iron transport [84]. HEPHL1 requires copper for stability [84], but is less investigated [41,43,64]. Molecular modeling on CP crystal structure indicates preserved HEPHL1 copper sites [84], and by analogy likely involved in copper mobilization. Recent proof is gained by a compound heterozygous patient with abnormal hair, joint laxity, and developmental delay (HJDD) [85]. Hair changes (pili torti and trichorrhexis nodosa) are similar to MNK, and muscular affection similar to SMAX3 [85]. LOX deficiency explains connective tissue involvement and indicates disturbed copper metabolism. HEPHL1 likely acts as redox chaperone in LOX cofactor formation and metalation. Lack of copper affects N-linked glycosylation sites, but none of the O-linked, indicating that copper is loaded before N-glycosylation [85].

4.2.4. Factor V+VIII Clotting Factors

Factor V+VIII clotting factors belong to blue copper oxidases with complex catalytic sites requiring ascorbate as cofactor [86,87]. Both are glycoproteins loaded by ATP7B before secretion to circulation. Within ER, FV and FVIII are guided by mannose binding lectin 1 (LMAN1) for trafficking and secretion [86,88,89]. LMAN1 receptor complex defects cause combined FV+VIII deficiency and mild coagulation disorder [90], underlining importance of mannose binding lectins in glycosylated protein trafficking. LMAN1 is a collectin requiring LOX activation, and in MNK low hepatic copper can affect maturation and trafficking leading to mild clotting deficiency [91].

4.3. Cytochrome c Oxidase

Cytochrome c oxidase (COX) uses copper and heme for reduction of oxygen, making mitochondrial copper and iron homeostasis indispensable for life. COX is the terminal oxidase of the electron transport chain comprising four complexes (numbered I, II, III, and IV), each with increasing reduction potential. The inner membrane contains respiratory complexes and copper chaperones; inter-membrane space (IMS) contains Cu/Zn superoxide dismutase 1 and soluble copper chaperones; matrix stores a mobilizable copper pool [92,93]. Copper delivery to COX requires an elaborate machinery of cytoplasmic guiding molecules, outer and inner membrane translocators, and embedded membrane chaperones [94]. Soluble and membrane anchored copper chaperones take different routes: (1) direct uptake via outer membrane into IMS using the redox import machinery [93], and (2) via matrix and redirection. Uptake of complex IV assembly factors is covered in reviews [94,95,96]. Matrix copper is routed to IMS for metalation and activation of enzymes and chaperones. Knowledge about copper exchange between mitochondrial compartments with different redox potentials is limited [97,98]. ATP7A dysfunction affects mitochondrial redox balance [99], and in MNK with low copper in liver and brain, COX deficiency becomes severe. COX is a multimeric complex containing two catalytic copper centers, CuA/heme and CuB, requiring assembly of numerous subunits. Formation of catalytic sites takes place within mitochondria and requires uptake of copper (and iron) into IMS from matrix pools, making insertion after biosynthesis unlikely. Assembly is complicated and assisted by several factors and copper chaperones [93,94,95,96]. The COX assembly will not be covered in detail. However, SCO1, SCO2, and COA6 are inner membrane embedded thiol-disulfide oxidorectases utilizing copper to modulate redox state, before delivery to CuA and CuB [100,101,102]. These copper chaperones are yet examples of redox-dependent copper transfer.
Respiratory chain secures cellular energy production, and disruption affects high energy-demanding tissues like CNS, liver, heart, and skeletal muscles. MNK shows numerous clinical signs of compromised mitochondrial activity [103,104,105]. Patients are hypotonic and floppy, often leading to suspicion of mitochondrial disturbance, and milder cases show myopathy. High lactate in blood or cerebrospinal fluid further strengthens a suspicion. Late disease stages develop deficiency of several respiratory chain complexes [105,106], and muscle ragged red fibers, a sign of mitochondrial dysfunction [103,106].

5. Copper-Catalyzed Cofactor Containing Enzymes

This enzyme group needs copper for cofactor activation. Each subclass has its unique activation, some are formed using copper during enzyme biogenesis, while others are attached to holoenzymes by a copper dependent process. SUMF1 dependent enzymes comprise a large new group of mainly lysosomal enzymes only recently joined with copper homeostasis

5.1. Copper Quinone Amine Oxidases

Copper quinone amine oxidases contain two subgroups, lysyl oxidases and copper amine oxidases, based on their specific internal cofactors that are formed post-translationally [107,108]. Copper-dependent monoamine oxidases are covered separately below, as they do not use an internal copper cofactor for catalysis. Notably, MAO A and B are not copper catalyzed.

5.1.1. Lysyl Oxidases (LOX)

Lysyl oxidases (LOX) comprise amine oxidases initiating extracellular matrix (ECM) formation by catalyzing oxidative deamination of an epsilon-amino group in lysine and hydroxylysine residues in first step of elastin and collagen cross-linking [108]. Importantly LOX cross-links a variety of triple helical proteins including collectins, significantly expanding LOX functions [109].
All members share a highly conserved catalytic site made up of three His [110] in close proximity to a cofactor created by copper catalyzed cross-linking of two internal amino acids [107,111]. Formation of the internal cofactor, LQT is described as an autocatalytic process, but new evidence points to a need for a redox process [85]. Copper LOX metalation occurs before N-glycosylation [85], and failure to acidify Golgi affects glycosylation resulting in cutis laxa [112,113].
Collagens (COL) and elastin (ELN) provide stability and connectivity among tissues and organs, and fiber building by trimerization and intra- and intermolecular cross-linking is crucial. Other components of ECM, mucopolysaccharides and mucolipids interact with fibers to form connective tissue. Lack of fiber tensile strength affects flexibility and integrity, and results in connective tissue disorders. Tissues affected are bones, tendons, ligaments, joints, muscles, blood vessels, heart, and eyes, with symptoms spanning from osteoporosis, loose joints, lung emphysema, aneurysms, glaucoma to pelvic organ prolapse or rupture [114]. New collectin symptoms are emerging. Numerous COL proteins exist, but type I is most abundant. Clinical defects span from osteogenesis imperfecta to osteoporosis and Ehlers-Danlos syndrome. ELN provides flexibility to fibers that rapidly expand and return to original shape. Deficiency results in joint laxity, wrinkling of skin, aneurysms and emphysema. Gastrointestinal and bladder diverticulae are common. Gly-X-Y triple repeats, hallmarks of COL are versatile and widespread in proteins adapted to a range of functions [115], and an increasing number of proteins with COL-like domains are identified [109]. Trimerization is crucial and needs LOX [115,116,117].
Amongst triple repeat proteins is complement Q1 (C1q), first member of the complement-cascade. Dysregulation of C1q is characterized by recurrent skin lesions, susceptibility to infections, increased risk of autoimmune disease, and chronic kidney disease [118,119]. C1q belongs to collectins [109,120] comprising a superfamily of lectins with a COL-like stretch fused with CRD. Collectins are present in plasma and on cell surfaces acting in first line of defense [109]. Lung surfactant collectins lubricate alveoli, besides being acute phase reactants.
A well described CRD receptor protein is mannose-binding lectin (lectin mannose binding 1) (LMAN1) that triggers the lectin pathway of complement activation. LMAN1 (often named ERGIC-53) is membrane bound and cycles between ER, ERGIC, and cis-Golgi [121,122,123]. LMAN1 and related proteins form complexes with lectin chaperones and is a rate-limiting step in maturation of secreted glycoproteins [89]. Compromised glycoprotein trafficking leads to incorrect localization, and mutations in LMAN1 receptor complex lead to combined deficiency of FV+VIII and hepatic accumulation of α-1-antitrypsin [90]. LMAN1 deficiency leads to susceptibility to meningitis and infections of upper respiratory tract, and as a trafficking factor for neuroreceptors will lead to CNS dysfunction [124,125]. More LMAN1 substrates are being identified also affecting immunobiology [123].
Numerous effector proteins exist, we include only a few here that are most important for MNK. Collagen-like tail of endplate acetylcholinesterase (COLQ) is acetylcholinesterase with a triple-helical membrane anchor to rapidly regulate muscle activation. If deficient, neuromuscular signaling causes muscle weakness [126]. Clinical signs are muscle fatiguability affecting limb muscles, ocular muscles (ptosis and ophthalmoplegia), and facial and mouth musculature (poor sucking and swallowing) as seen in MNK. Collagen and calcium binding EGF domains 1 (CCBE1) protein is important for lymphatic vessel formation, and deficiency results in lymphedema, and mutations in CCBE1 causes Hennekam syndrome [127]. MNK often encounter severe puffiness of face and feet, and a dough-like skin. Collectin defects are accompanied by susceptibility to infections [109], and distinctive facial features also seen in MNK, including widely spaced eyes (hypertelorism), narrowing of eye opening (blepharophimosis), droopy eyelids (ptosis), and high arched (cupid) eyebrows [128,129].
The LOX family is important in relation to MNK and contains five members, lysyl oxidase (LOX) and four lysyl oxidase-like (LOXL) enzymes working on different biological substrates [108,130,131]. LOX is secreted and cross-links ELN and COL. The proenzyme is activated attached to its extracellular substrates [131]. LOX plays a role in aortic wall formation, and deficiency predisposes to aortic aneurisms and dissections [132,133], a prevalent cause of death in MNK. Copper is loaded during Golgi passage, before final N-glycosylation, and a redox step is required [85]. LOX structure has been determined by homology modeling using LOXL2 [134] demonstrating copper coordination by three His and LQT oxygen [110] explaining a redox need in formation of the active center [85]. LOXL1 preferably cross-links ELN [135] and is linked to glaucoma, cataract [136,137], and lens zonule weakness eventually leading to lens subluxation. It is a major risk factor for pseudoexfoliation syndrome. Preferred substrate for LOXL2 is Type IV COL [138], a basement membrane component scaffolding other ECM molecules [130,139]. LOXL2 oxidizes histone and localizes in pericentromeric region [140]. Defects are associated with diseases of muscle, neural, ocular, cutaneous, vascular, lung and kidney tissues [141,142]. Substrates of LOXL3 are not clearly defined, but defects have been associated with early onset myopia [143] and Stickler syndrome that is a group of connective tissue defects with variable facial features, eye abnormalities, hearing loss, and joint problems [144]. LOXL3 localizes in nucleus and is involved in histone biology [145]. LOXL4 is expressed in cartilage and many tissues, the highest levels found in skeletal muscle, testis, and pancreas.
MNK shows numerous LOX and LOXL deficiency symptoms [146,147], and non-accidental injury (NAI) is often suspected [148,149,150], and an important differential diagnosis. LOX deficiency is also well established in OHS [1].

5.1.2. Copper-Containing Amine Oxidases (AOC)

Copper-containing amine oxidases (AOC) comprise a family of both diamine (DAO) and polyamine (PAO) oxidases. AOC participates together with numerous transporters and enzymes to precisely regulate polyamine pathways in CNS and periphery. Functions regulated are wakefulness, inflammation, and neurotransmitter release [108]. In addition, a catalytic role, copper is required for biogenesis of the internal topaquinone cofactor (TPQ) [108]. By analogy, a copper oxidase is likely required for formation of the active catalytic site where copper bridges three His and oxygen of TPQ. Histamine (monoamine), putrescine (diamine), spermidine (triamine), and spermine (tetramine) are ubiquitous AOC regulated amines, involved in proliferation, differentiation, and apoptosis, and in modulation of neurotransmitter receptors [151]. Polyamines play important roles in rapidly dividing cells like immune cells and enterocytes, and in regulation of membrane potentials in excitable tissues. Interactions with ligand-gated ion channels and tight-junctions are emerging as crucial polyamine regulated functions [151,152].
Polyamines are stable compounds present in mM amounts as free (minor), bound, and conjugated forms. Polyamine homeostasis is precisely regulated by de novo synthesis, extracellular catabolic control by AOC, intracellular regulatory feed-back loops, and membrane transfer by solute carriers (SLC) to balance intracellular and extracellular pools [153]. Physiological functions of charged polycations are not fully understood [153]. We focus on accumulation of specific polyamines, the result of deficient AOC catabolic control.
Histamine is the best studied amine, expressed at numerous sites including mast cells, gastrointestinal tract, and neurons [154]. Histamine regulates gastric acid secretion and CNS neurotransmission, in addition to a range of inflammatory reactions [155,156] mediated by specific histamine receptors [156,157].
Abnormal spermidine catabolism results in skin (ichthyosis), hair (alopecia), and eye (conjunctivitis) problems [158,159]. Allergic reactions (atopy), light intolerance (photophobia), and cornea inflammation (keratitis) may occur. High polyamine levels trigger persistent diarrhea with gastrointestinal polyps [154]. All these symptoms are found in MNK patients.
Humans have three AOC: AOC1, diamine oxidase, histaminase, or amiloride-binding protein 1 (ABP1), is mainly expressed in kidney, placenta, intestine, thymus, and seminal vesicles [108], and released at plasma membranes in response to external stimuli. AOC2, or retina-specific amine oxidase, is expressed on cell surfaces in many tissues with a particular high expression in retina [160]. AOC3, also named vascular adhesion protein 1 (VAP1), is widely distributed with highest expression in peripheral lymph nodes, hepatic endothelia, appendix, lung, and small intestine [108]. AOC3 has been implicated in lung inflammation, asthma, psoriasis, and vascular stroke [108]. Expression is also high in white fat tissue where it may be implicated in adipocyte differentiation and metabolism [161].

5.2. Formylglycine Activated Sulfatases

Sulfation/desulfation regulate numerous pathways, and sulfatases are responsible for break down and recycling of both complex sulfated sugars and hormones [162,163]. Sulfatases share a post-translationally formed internal cofactor, FGly essential for activity [164,165]. Cofactor generation requires sulfatase-modifying factor 1 (SUMF1) or formylglycine generating enzyme (FGE) [163,166]. SUMF1 oxidises cysteine in target enzymes using a highly conserved sequence, CXPSR [166,167], and recently copper was found to be required [167]. SUMF1 is an ER located soluble glycoprotein acting on native sulfatase polypeptides [168]. ER resident SUMF2 [169,170], a non-copper binding paralog acts as chaperone and retains SUMF1 by heterodimerization while activating sulfatases [171]. SUMF1 interacts with numerous trafficking factors including LMAN1, and lack of activation and trafficking leads to proteasomal degradation of SUMF1 [172]. Sulfatases localize to subcellular sites such as lysosomes, Golgi, and ER [170], where they break-down complex mucopolysaccharides, mucolipids, and steroid hormones. Lysosomal glycosaminoglycan (GAG) sulfatases comprise a major group [173]. GAGs are complex sugar polymers and important components of bone and cartilage, joint lubricants, and cell surface initiating growth factor activity and first line of defense against microorganisms. Recycling of GAGs starts by removal of sulfated groups and defective recycling results in GAG accumulation. Deficiencies present as mimicry of mucopolysaccharidoses (MPS) and mucolipidoses (MLP) affecting multiple organ systems [162,170,173]. Sulfatation/desulfatation are crucial for cartilage formation, and defects are often accompanied by bone dysplasias [173].
Most steroids, e.g., cholesterol, pregnenolone, and estrone, are sulfated after biosynthesis [162], and sulfatation is vital for endocrine function. Cholesterol is crucial for neurotransmission, myelination, and synaptogenesis [174], and desulfatation provides a copper link. Dysregulation is associated with numerous pathologies, including faulty regulation of GABA receptor function [175,176]. Niemann-Pick C disease may be accompanied by copper disturbance likely secondary to poor steroid sulfatase activity and disrupted trafficking of cholesterol [177].
Combined impairment of all sulfatases, multiple sulfatase deficiency (MSD), are clinically heterogeneous disorders caused by mutations in SUMF1 or SUMF2 [169,170]. Symptoms present features of metachromatic leukodystrophy, mucopolysaccharidosis, chondrodysplasia punctata, hydrocephalus, ichthyosis, neurological deterioration, and developmental delay.
ATP7A-related disturbances may mimic MSD and present with overlapping clinical features from a complex interplay between SUMF1 and the LOX family. Sulfated molecules build up in lysosomes, resulting in necrosis and metachromasia, a sign noted early in MNK [178], but forgotten when the copper disturbance was discovered. Morphologic changes with vacuoles in myeloid cells, termed Alder Reilly anomaly are seen in patients with mucopolysaccharidoses (MPS) and have also been reported in MNK [179,180]. Skin problems in MNK may be related to deficient steroid sulfatase (ichthyosis) [181,182] also affecting keratinocyte biogenesis and hair development [183]. Build-up of cholesterol sulfate in the outermost layer of epidermis causes hyperkeratosis with scaling [184].

6. Copper-Dependent Mono-Amine Oxidases

Copper monooxygenases catalyze reactions in catecholamine and hormone pathways. The group consists of four enzymes that are free or membrane attached within vesicles of same embryonic origin: adrenal chromaffin vesicles (DBH), synaptic vesicles of the sympathetic nervous system (DBH), secretory vesicles of the pituitary gland (PAM), and melanocytes in periphery and CNS (TYR). Enzymes travel to their final destination, but trafficking is not completely understood and depend on metalation and N-glycosylation. Crystal structure of DBH shows two copper sites, one (CuH) coordinated by three His, the other (CuM) by two His and one Met [185]. Topology is similar to PAM, and also shows likeness to TYR [185,186]. All sites have similar copper avidity [28].

6.1. Dopamine β-Hydroxylase (DBH)

Dopamine β-hydroxylase (DBH) is an ascorbate-dependent monooxygenase converting dopamine (DA) to norepinephrine (NE) [187,188]. DBH localizes in synaptic vesicles in noradrenergic and adrenergic nerve terminals of central and peripheral nervous system, as well as adrenal medulla [189]. DBH is targeted to secretory granules by ER glycosylation [185].
DBH contains three N-glycosylation sites [185], and may show trafficking problems [190], and misfolding is suggested to cause DBH deficiency [190]. ATP7A supplies copper to DBH both centrally and in the periphery [191,192,193] and is needed during formation and maturation of the holoenzyme, though copper can likely be loaded later. Met-His-rich lumenal loop of ATP7A can experimentally transfer copper to DBH [28]. NE controls mood, attention, and overall arousal, as well as stress, learning, and memory [185], and the adrenal system is important in maintaining blood pressure, glucose, and sodium levels [194]. Congenital NE deficiency shows profound autonomic failure [188], and perinatal period may be complicated by vomiting, dehydration, hypotension, hypothermia, and severe hypoglycemia all seen in early MNK. Later symptoms are dizziness upon standing (orthostatic hypotension), blurred vision, and difficulty in exercising. Other symptoms are droopy eyelids (ptosis), nasal congestion, muscle pain, and weakness, symptoms well recognized in MNK. DA/NE ratio is increased in plasma and CSF, and dopaminergic imbalance is an early discriminatory marker for MNK [195], and milder forms also show abnormal values [2].

6.2. Peptidyl α-Amidating Enzyme (PAM)

Peptidyl α-amidating enzyme (PAM) activates a vast amount of neuroendocrine hormones involved in regulation of numerous processes. PAM is a bifunctional enzyme, consisting of two distinct catalytic domains working sequentially, peptidylglycine α-hydroxylating monooxygenase (PHM) and peptidyl–α-hydroxyglycine α-amidating lyase (PAL). Copper containing PHM catalyzes hydroxylation of a glycine, subsequently cleaved by PAL to generate C-terminal amidation in activated peptide hormones. PHM has copper-binding sites similar to DBH and also requires ascorbate as cofactor [185,186]. Lack of metalation does not alter passage through secretory pathway, and the apoenzyme is not degraded [196], though not directed to correct vesicular location [186]. Copper required for enzyme activity is not tightly bound [7] and can be lost, but secreted apoenzyme can be activated [197]. This likely also apply for trafficking and metalation of related enzymes, DBH and TYR. The first luminal loop of ATP7A involved in release of copper contains an amino acid stretch rich in His and Met acting as potential copper donor for metalation of PAM in secretory pathway [186]. Functionally PAM and DBH overlap, and neuropeptides and neurotransmitters participate in a large number of processes related to feeding and body weight, fluid balance, pain, anxiety, memory, circadian rhythms, and reward [186,198]. PAM is essential for activation of numerous neuroendocrine peptide hormones such as cholecystokinin, gastrin, vasoactive intestinal peptide, thyrotropin-releasing hormone, calcitonin, corticotropin-releasing hormone, and vasopressin [186].
Biological significance of PAM is not fully understood, but deficiency results in widespread effects. Brindled mice, a genetic model of ATP7A-related copper disturbances, fail to produce normal levels of α-amidated peptides [198,199]. PAM deficient mice show CNS problems, e.g., impaired vasoconstriction and thermoregulation, increased seizure susceptibility, anxiety, and increased response to noise [198].

6.3. Monooxygenase, DBH-Like 1 (MOXD1)

Monooxygenase, DBH-like 1 (MOXD1) is structurally similar to other ascorbate requiring copper-containing monooxygenases, but with unknown substrate. MOXD1 lacks signal sequence and localizes throughout ER in both endocrine and non-endocrine cells [200]. MOXD1 is membrane-associated and oligomerize. MOXD1 is predicted to hydroxylate a substrate in ER, and possibly acts as enzyme chaperone for DBH [200].

6.4. Tyrosinase (TYR)

Tyrosinase (TYR) catalyzes the first two steps in melanogenesis from tyrosine to DOPA and to dopaquinone. Tyrosine oxidation is rate-limiting followed by ER polymerization reactions [201,202] catalyzed by two members of tyrosinase-related proteins TYRP1 and TYRP2 [203].
TYR is membrane anchored and possesses two copper centers resembling DBH though entirely made up of His [185,204]. Copper is acquired during maturation in secretory pathway, but apo-TYR can be activated later by addition of copper [205,206]. TYR localizes to specialized endosomes termed melanosomes and undergoes maturation and sorting before reaching integration site [207,208]. Intracellular sorting and polymerization steps from ER through Golgi to melanosomes is tightly regulated including metalation and N-glycosylation [203,209,210]. During sorting in ER, TYR interacts with lectins normally associated with LMAN1 [207,209]. Metalation likely occurs in ER before action of TYRP1 and TYRP2, but can take place later in melanosomes, and TYR becomes fully functional only at its final destination [210,211]. TYRP1 and TYRP2 belong to the same protein family and have similar metal binding sites though using zinc. TYR substrates play a conformational role as molecular chaperones to enhance folding and ERGIC trafficking [208]. The Met-His-rich first luminal loop of ATP7A possibly metalates TYR [28]. TYR may lose copper during passage of acidic TGN but is reloaded in melanosomes with neutral pH [210,212].
Melanosomes originate from distinct, though related, embryonic stem cells: (1) neural tube derived retinal pigment epithelium and pineal gland melanocytes; (2) neural crest derived melanocytes of inner ear, skin, hair-bulbs, and iris [213]. Highest TYR expression is in pigment epithelium of retina. Skin melanosomes are transferred to keratinocytes where melanins protect against UV sun radiation [214]. Melanins are negatively charged, polymerized and hydrophobic pigments working as capacitor to absorb and dissipate energy to neutralize radiation. In case of high energy absorption, output occurs as heat and reactive oxygen species (ROS), eventually resulting in sun burn and necrosis. Complex neuromelanins are synthesized mainly in dopaminergic neurons of substantia nigra and noradrenergic neurons of locus coeruleus [215]. Midbrain catecholaminergic neurons of basal ganglia network are crucial for brain cognitive functions. Biosynthesis and regulation of neuromelanins are poorly understood [216,217] as is their role in smell, vision, and hearing. Deficient development of inner ear melanocytes causes deafness [218,219]. TYR mutations result in hypopigmentation disorders and sensitivity to UV radiation, visual problems like nystagmus, strabismus, and reduced visual acuity with photophobia [220]. Transduction overload may lead to local oxidative stress and accumulation of waste products in central and peripheral ganglions and increased risk of melanoma [217]. Pigment and cell debris accumulation in CNS may increase susceptibility to Parkinson [217].
In MNK visual problems are early onset nystagmus, iris trans-luminescence, hypo-pigmented fundus, and reduced visual acuity [221]. Hearing may be impaired, but often not investigated. In accord with above, copper replacement therapy in MNK shows darkening of hair and skin [206].

7. Copper/Zinc-Containing Superoxide Dismutases (Cu/Zn-SODs)

Superoxides are products of normal aerobic metabolism and crucial in oxidative burst of innate immune responses [222], but in need of strict control. Superoxide dismutase (SOD) disproportionate the reactive radicals into molecular oxygen and less reactive hydrogen peroxide. Uncontrolled, ROS will attack unsaturated fatty acids, and SOD is of particular importance for a healthy brain, and of the most abundant enzymes underlining importance of ROS control. SOD1 is compartmentalized into distinct cellular and minor extracellular pools. SOD3 is attached to extracellular matrix, and often named extracellular SOD (EC-SOD). SOD1 activity is copper regulated at protein level, while SOD3 activity is copper regulated at gene level. Amyloid-β precursor protein (APP) family consists of Cu/Zn proteins with a SOD-like structure and possible dismutase activity [223] and is copper regulated through Cu-CCS activated cleavage of β-secretase 1 (BACE1). A manganese form (SOD2) in mitochondrial matrix is interconnected with IMS-SOD1 [224] and SOD3 [225].

7.1. Superoxide Dismutase 1 (SOD1)

Superoxide dismutase 1 (SOD1) is the master SOD and sole cytosolic and peroxisomal cuproenzyme. SOD1 mainly localizes in cytosol, an almost equal fraction in peroxisomes [226], and minor pools in mitochondrial intermembrane space (IMS), and nucleus. SOD1 comprise a large copper pool, earlier viewed as copper buffer [227], substantiated by labile metal binding by a cluster of four imidazole groups [11]. Some cell types secrete SOD1 [228]. SOD1 is unusual by having a labile copper site in cytoplasm abundant in GSH, and likely needs shielding by vesicular structures [229]. Copper chaperone for SOD1 (CCS) participates in maturation and activation of SOD1 at all subcellular locations. CCS is a member of the Cu/Zn-SOD family and acts as an enzyme chaperone to catalyze an intramolecular disulfide bond, stabilizing correct SOD1 conformation for incorporation of copper and zinc. CCS also functions as molecular chaperone and contains three domains having different roles: N-terminus possesses a copper binding site (MXCXXC), similar to ATOX1 [230], also with potential allosteric role in copper activation. The homologous middle part interacts with SOD1, and C-terminus contains a copper catalytic CXC site needed for intramolecular S-S bridge formation [231,232]. Nascent SOD1 and CCS polypeptides devoid of metal enter IMS individually and with essential sulfides reduced while traversing outer mitochondrial membrane [93,233]. In IMS, SOD1 meets CCS, and is folded and activated as in cytosol, hereby retained in IMS as functional enzyme. SOD1 and CCS are both taken up by the CHCHD4 (~MIA40) redox import machinery [234].
Peroxisomes enclosed by a single lipid bilayer use special import of membrane proteins and matrix enzymes [235]. Most contains a peroxisomal targeting signal (PTS) and are taken up via peroxin (PEX) membrane receptors [236,237] and delivered through direct contact between ER and peroxisomes. Folded, co-factor bound, and oligomeric proteins can be imported [238]. A major SOD1 route through ER has been discovered, securing high peroxisomal matrix content [239]. SOD1 does not contain PTS and is piggy-backed into peroxisomes by its chaperone [237,239]. CCS-PTS is in ER recognized by PEX5 receptor, shuttling CCS-SOD1 into peroxisomal matrix [235]. SOD1 rapidly enters nucleus in response to increased H2O2 levels and is potentially piggy-backed via ER by CCS. Peroxisomes are present in all tissues catalyzing a wide range of anabolic and catabolic reactions. SOD1 generates H2O2, and catalase uses H2O2 to oxidize substrates. SOD1 dysfunction leads to ROS accumulation that eventually damage the peroxisomal membrane, and release catalase to cytosol [240]. Severe pathologies result from peroxisomal dysfunction showing multi-systemic symptoms referred to as peroxisome biogenesis disorders (PBD). Neurological dysfunction is prominent usually accompanied by brain malformations, myelin abnormalities, and neuronal degeneration [241]. Systemic manifestations often include dysmorphic features, liver dysfunction, and skeletal abnormalities [241].
In MNK brain, both CCS and SOD1 polypeptides are taken up into mitochondrial IMS, but SOD1 is not properly folded and activated due to lack of copper. ROS are expectedly high, and matrix SOD2 induced as compensation [224]. The ER-peroxisomal route is also compromised creating a deficit of peroxisomal matrix SOD1 and enhanced peroxisomal stress in turn affecting nerve development. Still CCS accumulates [224,242] indicating faulty heterodimerization when copper is low. Deficiencies affect cerebellar maturation and axonal integrity, and lead to Purkinje cell pathologies with “weeping willow”, a well-recognized sign in MNK [242,243]. Low hepatic copper results in low SOD1, and oxidative stress plays a role in the pathogenesis of steatosis.
If nascent SOD1 is not correctly processed, it will remain inactive, potentially misfold, dimerize or tetramerize, as is the case in some neurodegenerative diseases [244]. Genetic disturbances of SOD1 lead to motor neuron disease, amyotrophic lateral sclerosis (ALS). Most SOD1 mutations affects heterodimerization and piggy-backing into peroxisomes [245]. Like MNK, lack of peroxisomal uptake of SOD1 will in ALS lead to oxidative stress and development of varying motor neuron affection as part of the PBD spectrum.

7.2. Superoxide Dismutase 3 (SOD3)

Extracellular superoxide dismutase (SOD3) is anchored to heparan sulfate in ECM [246]. SOD3 is structurally closely related to SOD1 and also contains copper in its catalytic center and zinc to stabilize structure. The central part of SOD3 is homologous to SOD1, the metal binding sites preserved and with similar copper avidity, but structures vary at ends. SOD3 contains a signal peptide plus three N-glycosylation sites for GAG guidance [247]. The enzyme is copper loaded in secretory pathway, but no specific copper chaperone has been identified. ATOX1 regulates protein expression through copper dependent binding to SOD3 promoter [248]. C-terminus contains a heparan binding domain securing attachment to ECM [246]. After secretion SOD3 forms tetramers stabilized by intermolecular disulfide bonds. SOD3 is secreted by fibroblasts and glial cells and protects cell membranes against ROS; about 1% is free in plasma, lymph, and cerebrospinal fluid [249]. SOD3 levels are high in vasculature, heart, lungs, kidney, and placenta [250]. Low SOD3 activity is linked to lung disease such as acute respiratory distress syndrome or chronic obstructive pulmonary disease [251] and deficiency may result in angiopathy.

8. Cu/Zn-SOD-Related Proteins Regulated by β-Secretase 1 (BACE1)

Amyloid-β precursor protein (APP), and amyloid-like proteins APLP1 and APLP2 contain a dismutase fold resembling Cu/Zn-dismutases and are regulated by copper through protease cleavage [11,224]. They bind copper and zinc primarily through His coordination [252] but an enzymatic role has not been established, though APP redox capacity has been demonstrated [253,254]. APP and its processed forms appear to have a growth-factor-like role and promotes neuronal proliferation and division [255]. Thus, the APP family is important for synaptic development and plasticity of central and peripheral nervous systems [256]. We will point to the relationship between the APP family and the Cu/Zn-SOD family to emphasize remote regulation by CCS-Cu.
β-secretase 1 (BACE1) is a membrane-bound protease, catalyzing first step of extracellular release of soluble amyloid β peptide (Abeta) from APP. BACE1 is rate-limiting in neuronal Abeta generation and also cleaves numerous other substrates important in formation of myelin. BACE1 contains a CCS-Cu regulatory site spatially separated from the protease site [5,11]. N-terminal CCS-MXCXXC binds to a cysteine rich area in C-terminal cytoplasmic tail of BACE1 regulating numerous brain functions including PAM [257]. BACE1 is expressed at high levels in brain and pancreas. Expression is highest in substantia nigra, locus coeruleus and medulla oblongata [258]. Abrogated cleavage in BACE1 knockout mice shows a role in neuronal migration, axonal growth, and muscle spindle function [259]. BACE1 is N-glycosylated [260] implying poor ERGIC trafficking in addition to poor protease activity secondary to low brain copper in MNK.

9. Conclusions

The main objective of this review is tying enzymes, substrates, and key symptoms together in a unified hypothesis to explain Menkes disease symptoms and pathologies (Table 3). We also wish to shed light on crucial steps in biogenesis of copper-dependent enzymes (dysfunctional in MNK) by focusing on metalation sites in cells, metal chaperoning and trafficking of enzymes in the secretory pathway.
ATP7A disturbances result in complicated copper disorders starting by poor uptake at intestinal brush border, aggravated by poor release from enterocytes, further affecting all barriers in the body, underlining that the basic defect is not a simple copper insufficiency. Defects in reduction (STEAP) before cellular uptake and in oxidation (HEPH) before release contribute to a complex copper transport defect resulting in complex clinical traits. Intracellular organelle deficiencies develop, combined with copper accumulation in unavailable pools. Copper pumping into secretory pathway and enzyme metalation are clinically significant, and ERGIC enzyme trafficking is also emerging as a copper regulated step (LOX). MNK diagnosis is often missed until hair changes are obvious, and the delay may leave many undiagnosed cases. To improve diagnostic awareness, focus should be shifted from hair as the main diagnostic pointer to more subtle symptoms. We found no evidence of a copper specific sulfhydryl oxidase, and hair and skin changes likely result from combined lack of steroid sulfatase (SUMF1), copper amine oxidase (AOC), and defective mitochondrial Fe-S biogenesis. SUMF1 is a new player in Menkes disease linking faulty cholesterol biology to the clinical picture and a whole new group of GAG sulfatases, which may lead to mimicry of lysosomal storage disorders (Table 3).
Symptoms secondary to LOX dysfunction have been expanded and shed light on their role in activation of receptor and adapter collectin molecules. Though important, it is an overlooked component of Menkes disease pathology. In liver, LMAN1 deficiency affects coagulation factors V+VIII and alpha-1-antitrypsin, and in brain leads to poor trafficking of numerous neuroreceptors explaining nervous symptoms in MNK. Other receptors with a collagen-like stretch, COLQ and CCBE1, explain muscle weakness and lymphedema. Cq1 deficiency add problems with innate immunity, and lung surfactant defects. Unexpectedly the peroxisomal SOD1 pool requires ER for metalation, and Zellweger-like symptoms are becoming part of the MNK symptom spectrum. Interestingly, motor neuron disease is a characteristic of the mildest disease form, SMAX3.
Trafficking and post-translational modifications of copper enzymes, including metalation begin in the endoplasmic reticulum (ER) and continues in Golgi before proteins are sorted and sent to their final destinations. Sugar tags guide enzymes during folding, proof-reading, refolding, and holoenzyme trafficking [261]. In ER nascent polypeptides are core glycosylated, and the added sugar tags are used for cargo receptor recognition by LMAN1 and other lectins. Adaptor sugar recognition is important for correct folding and trafficking, and GAG defects lead to multiple tissue and organ failures as well as abnormal physiognomy. Proteins with a CRD domain constitute a distinct class of adaptor molecules of which collectins [109] require LOX for correct conformation and stability.
At present, the extent of glycosylation and trafficking defects in Menkes disease is unclear, and coppers significance for sugar sorting is an emerging field. LMAN1 is one of several homologous mannose binding adapter molecules securing protein trafficking in the secretory pathway. Further glycosylation modifications occur in the Golgi complex where an array of enzymes modifies the sugar tags for their final destination but may require metalation to expose N-glycosylation sites correctly [261]. Lack of copper can result in distorted conformation and lead to normally unexposed N-glycosylation sites being exposed or the opposite, resulting in integration at wrong membrane sites [261].
Copper metalation is most often cited as taking place in TGN, but we found clear evidence in the literature of metalation in ER. Possibly ATP7A delivers copper in ER, in ERGIC, and in Golgi. SUMF1 is resident and metalated in ER and requires an ER-resident homologue devoid of copper, SUMF2 as molecular chaperone. Strong indication exists that TYR and DBH are metalated in ER, also making ER-metalation of PHM likely. TYR sites have low avidity and if pH is low, often loses copper, but is reloaded in melanocytes with a neutral pH. Potentially ATP7A also provides copper here. TYR uses related molecules, TYRP1 and TYRP2 as molecular chaperones. TYRP1 and TYRP2 contain zinc and are both ER located. DBH is suggested to use the ER-resident homologue, MOX1D as molecular chaperone.
Blue copper oxidases (CP, HEPH, and HEPHL1) appear to be metalated in cis-Golgi. Cofactor formation of LOX and AOC happens by use of a redox process, before N-glycosylation in Golgi, though the process is normally cited as autocatalytic. However, required reactions will likely not rely on chance, but is facilitated by an enzyme reaction. HEPHL1 is needed for metalation of LOX, and AOC likely use the same or a similar redox chaperone.
SOD1 is metalated at several cellular sites and depends on ER for metalation of the peroxisomal pool. CCS does not load SOD1 but is needed as redox chaperone to form S-S bridges stabilizing the conformation for proper metalation and subsequent piggy-backing of the CCS-SOD1 complex to peroxisomes. CCS provides allosteric regulation of SOD1 and BACE1, similarly to ATOX1 that allosterically regulates MBDs to initiate ATP7A/B pump activity.
ATP7A contains a Golgi localization signal and locates in ER when the signal is removed by alternative splicing. The first lumenal loop may help retain the protein in ER by binding of copper to Met-His-rich sequences similar to calcium ATPases using corresponding sites for calcium regulated ER retention.Metalation of DBH, PAM, and TYR may be facilitated by the Met-His-rich lumenal loop of ATP7A.
At experimental tissue culture conditions, ATP7A is found in TGN. However, most tissue culture experiments use fibroblasts, and the principal enzyme in this cell type is LOX, which is metalated in the late secretory pathway. Fet3 models will misinterpret ER activity as it is a CP/HEPH homologue metalated in Golgi. Thus, experiments using tissue culture may not represent the full picture of what takes place in vivo. We hypothesize that if ER metalation is diminished, all enzymes including downstream metalated enzymes may be affected leading to the severest phenotype. Milder phenotypes may preserve ER metalation of enzymes but show Golgi metalation problems. However, enzymes with low copper avidity may lose the metal during Golgi passage, and the enzyme may integrate at a faulty site resulting in deficient function, as may be the case for, e.g., DBH and TYR. Notably, all ATP7A-related phenotypes except SMAX3 show pale skin color and dysautonomia.

10. Future Directions

Numerous copper chaperones and adaptor molecules regulate copper-dependent enzyme passage in the secretory pathway, but the specific guiding molecules are only known for a fraction of copper enzymes. We hypothesize that more copper-dependent enzymes need specific copper chaperones for metal activation, and chaperoning roles may emerge for copper binding proteins for which today there is no known function. For example, the APP family may act as redox-active copper chaperones similarly to CCS. More copper-dependent enzyme reactions are likely to be unraveled, e.g., mitochondrial enzymes controlled by lipoic acid may also depend on copper. Finally, the iron–copper connection needs to be further explored on molecular, cellular, and organelle levels. Despite the new cellular/molecular connections outlined here for copper-dependent processes, the Menkes disease enzyme puzzle, linking consequences of ATP7A dysfunction in cells and tissue to MNK patients’ clinical symptoms, is not yet complete.

Author Contributions

N.H. conceived the idea and wrote the first draft. N.H. and P.W.-S. edited the text together. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ABP1Amiloride-binding protein 1
ALSAmyotrophic lateral sclerosis
AMDAge-related macular degeneration
AOCCopper-containing amine oxidases
APPAmyloid-β precursor protein
APLPAmyloid-like protein
ARSArylsulfatase
ATOX1Antioxidant 1 copper chaperone
ATPAdenosine triphosphate
ATP7ACopper-transporting ATPases, α
ATP7BCopper-transporting ATPases, β
BACE1β-secretase 1
BBBBlood–brain barrier
CCysteine
C1qComplement Q1
CCBE1 Collagen and calcium binding EGF domains 1
CCSCopper chaperone for SOD1
CHCHDCoiled-coil-coiled-coil domain
CNSCentral Nervous System
COA6COX Assembly Factor 6
COL Collagen
COLQ Collagen-like tail of endplate acetylcholinesterase
COXCytochrom c Oxidase
CPCeruloplasmin
CRD Carbohydrate-recognition domain
CSF Cerebrospinal Fluid
Cu Copper
CysCysteine
CYBRD1Cytochrome b reductase 1
DADopamine
DAODiamine Oxidase
DBHDopamine β-hydroxylase
DOPADihydroxyphenylalanine
ECMExtracellular Matrix
EC Extracellular
ELNElastin
EREndoplasmic reticulum
ERGICER–Golgi intermediate compartment
FADFlavin adenine dinucleotide
FV+VIII Clotting factors V and VIII
Fe-SIron sulfur site
FGEFormylglycine generating enzyme
FGlyFormylglycine
GGlycine
GABAGamma aminobutyric acid
GAGGlycosaminoglycan
GALNSGalactosamine-6-sulfate sulfatase
GlyGlycine
GMXCXXCAmino acid sequence of ATP7A/B MBD
GNSN-acetylglucosamine-6-sulfatase
GPIGlycosylphosphatidylinositol
GSHGlutathione
HEPHHephaestin
HEPHL1Hephaestin-like protein 1
HisHistidine
H2O2Hydrogen peroxide
IDSIdurunate 2-sulfatase
IMMInner mitochondrial membrane
IMSInter-membrane space
LMANLectin mannose binding
LOX Lysyl oxidase
LOXL Lysyl oxidase-like
LTQLysine tyrosylquinone
MMethionine
MNKMenkes disease
MBDMetal binding domain
MXCXXC Amino acid sequence of CCS-MBD
MetMethionine
MIA40Mitochondrial IMS assembly 40
MNDMotor neuron disease
MLPMucolipidose
MPSMucopolysaccharidose
MSDMultiple sulfatase deficiency
MOXD1Monooxygenase, DBH-Like 1
NAD(P)HNicotinamide adenine dinucleotide phosphate
NENorepinephrine
NLSNuclear localization sequence
OHSOccipital horn syndrome
OMIMOnline mendelian inheritance in man
PALPeptidyl–α-hydroxyglycine α-amidating lyase
PAMPeptidyl α-amidating enzyme
PAOPolyamine oxidase
PEXPeroxin
PHMPeptidylglycine α-hydroxylating monooxygenase
PTSPeroxisomal targeting signal
ROSReactive oxygen species
SCOSynthesis of COX
SMAX3X-linked distal spinal muscular atrophy 3
SODSuperoxide dismutase
SPSecretory pathway
S-SDisulfide bridge
STEAPSix-transmembrane epithelial antigen of prostate
STSSteroid sulfatase
SUMFSulfatase-modifying factor
TGN:Trans Golgi Network
TMTransmembrane
TPQTrihydroxyphenylalanine quinone
TYRTyrosinase
TYRPTyrosinase related protein
VAP-1Vascular adhesion protein 1
PBDPeroxisome biogenesis disorders
ZnZinc

References

  1. Horn, N.; Tümer, Z. Menkes Disease and the Occipital Horn Syndrome. In Connective Tissue and Its Heritable Disorders, 2nd ed.; Royce, P., Steinmann, B., Eds.; Wiley-Liss: New York, NY, USA, 2002; Volume 14, pp. 651–685. [Google Scholar]
  2. Kaler, S.G. ATP7A-Related Copper Transport Disorders. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Mirzaa, G., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 2003. Available online: https://www.ncbi.nlm.nih.gov/books/ (accessed on 4 January 2021).
  3. Zlatic, S.; Comstra, H.S.; Gokhale, A.; Petris, M.J.; Faundez, V. Molecular basis of neurodegeneration and neurodevelopmental defects in Menkes disease. Neurobiol. Dis. 2015, 81, 154–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Horn, N.; Møller, L.B.; Nurchi, V.M.; Aaseth, J. Chelating principles in Menkes and Wilson diseases: Choosing the right compounds in the right combinations at the right time. J. Inorg. Biochem. 2019, 190, 98–112. [Google Scholar] [CrossRef] [PubMed]
  5. Angeletti, B.; Waldron, K.J.; Freeman, K.B.; Bawagan, H.; Hussain, I.; Miller, C.C.; Lau, K.F.; Tennant, M.E.; Dennison, C.; Robinson, N.J.; et al. BACE1 cytoplasmic domain interacts with the copper chaperone for superoxide dismutase-1 and binds copper. J. Biol. Chem. 2005, 280, 17930–17937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Solomon, E.I.; Heppner, D.E.; Johnston, E.M.; Ginsbach, J.W.; Cirera, J.; Qayyum, M.; Kieber-Emmons, M.T.; Kjaergaard, C.H.; Hadt, R.G.; Tian, L. Copper active sites in biology. Chem. Rev. 2014, 114, 3659–3853. [Google Scholar] [CrossRef] [Green Version]
  7. Maheshwari, S.; Shimokawa, C.; Rudzka, K.; Kline, C.D.; Eipper, B.A.; Mains, R.E.; Gabelli, S.B.; Blackburn, N.; Amzel, L.M. Effects of copper occupancy on the conformational landscape of peptidylglycine α-hydroxylating monooxygenase. Commun. Biol. 2018, 1, 74. [Google Scholar] [CrossRef] [PubMed]
  8. Hatori, Y.; Yan, Y.; Schmidt, K.; Furukawa, E.; Hasan, N.M.; Yang, N.; Liu, C.N.; Sockanathan, S.; Lutsenko, S. Neuronal differentiation is associated with a redox-regulated increase of copper flow to the secretory pathway. Nat. Commun. 2016, 7, 10640. [Google Scholar] [CrossRef] [Green Version]
  9. Kellokumpu, S. Golgi pH, Ion and Redox Homeostasis: How Much Do They Really Matter? Front. Cell Dev. Biol. 2019, 7, 93. [Google Scholar] [CrossRef] [Green Version]
  10. Crisponi, G.; Nurchi, V.M.; Fanni, D.; Gerosa, C.; Nemolato, S.; Faa, G. Copper-related diseases: From chemistry to molecular pathology. Coord. Chem. Rev. 2010, 254, 876–889. [Google Scholar] [CrossRef]
  11. Öhrvik, H.; Aaseth, J.; Horn, N. Orchestration of Dynamic Copper Navigation—New and Missing Pieces. Metallomics 2017, 9, 1204–1229. [Google Scholar] [CrossRef]
  12. Shanbhag, V.C.; Gudekar, N.; Jasmer, K.; Papageorgiou, C.; Singh, K.; Petris, M.J. Copper metabolism as a unique vulnerability in cancer. Biochim. Biophys. Acta. Mol. Cell. Res. 2021, 1868, 118893. [Google Scholar] [CrossRef]
  13. Lutsenko, S.; Barnes, N.L.; Bartee, M.Y.; Dmitriev, O.Y. Function and regulation of human copper-transporting ATPases. Physiol. Rev. 2007, 87, 1011–1046. [Google Scholar] [CrossRef]
  14. Linz, R.; Lutsenko, S. Copper-transporting ATPases ATP7A and ATP7B: Cousins, not twins. J. Bioenerg. Biomembr. 2007, 39, 403–407. [Google Scholar] [CrossRef] [PubMed]
  15. Braiterman, L.; Nyasae, L.; Guo, Y.; Bustos, R.; Lutsenko, S.; Hubbard, A. Apical targeting and Golgi retention signals reside within a 9-amino acid sequence in the copper-ATPase, ATP7B. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 296, G433–G444. [Google Scholar] [CrossRef] [PubMed]
  16. Greenough, M.; Pase, L.; Voskoboinik, I.; Petris, M.J.; O’Brien, A.W.; Camakaris, J. Signals regulating trafficking of Menkes (MNK.; ATP7A) copper-translocating P-type ATPase in polarized MDCK cells. Am. J. Physiol. Cell Physiol. 2004, 287, C1463–C1471. [Google Scholar] [CrossRef] [PubMed]
  17. Tadini-Buoninsegni, F.; Smeazzetto, S. Mechanisms of charge transfer in human copper ATPases ATP7A and ATP7B. IUBMB Life 2017, 69, 218–225. [Google Scholar] [CrossRef] [Green Version]
  18. Yu, C.H.; Dolgova, N.V.; Dmitriev, O.Y. Dynamics of the metal binding domains and regulation of the human copper transporters ATP7B and ATP7A. IUBMB Life 2017, 69, 226–235. [Google Scholar] [CrossRef]
  19. Banci, L.; Bertini, I.; Cantini, F.; Rosenzweig, A.C.; Yatsunyk, L.A. Metal Binding Domains 3 and 4 of the Wilson Disease Protein: Solution Structure and Interaction with the Copper(I) Chaperone HAH1. Biochemistry 2008, 47, 7423–7429. [Google Scholar] [CrossRef] [Green Version]
  20. Francis, M.J.; Jones, E.E.; Levy, E.R.; Ponnambalam, S.; Chelly, J.; Monaco, A.P. A Golgi localization signal identified in the Menkes recombinant protein. Hum. Mol. Genet. 1998, 7, 1245–1252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Ariöz, C.; Wittung-Stafshede, P. Folding of copper proteins: Role of the metal? Q. Rev. Biophys. 2018, 51, e4. [Google Scholar] [CrossRef] [PubMed]
  22. Singleton, W.C.; McInnes, K.T.; Cater, M.A.; Winnall, W.R.; McKirdy, R.; Yu, Y.; Taylor, P.E.; Ke, B.X.; Richardson, D.R.; Mercer, J.F.; et al. Role of glutaredoxin1 and glutathione in regulating the activity of the copper-transporting P-type ATPases, ATP7A and ATP7B. J. Biol. Chem. 2010, 285, 27111–27121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Grek, C.L.; Zhang, J.; Manevich, Y.; Townsend, D.M.; Tew, K.D. Causes and consequences of cysteine S-glutathionylation. J. Biol. Chem. 2013, 288, 26497–26504. [Google Scholar] [CrossRef] [Green Version]
  24. Qi, M.; Byers, P.H. Constitutive skipping of alternatively spliced exon 10 in the ATP7A gene abolishes Golgi localization of the menkes protein and produces the occipital horn syndrome. Hum. Mol. Genet. 1998, 7, 465–469. [Google Scholar] [CrossRef] [Green Version]
  25. Barry, A.N.; Otoikhian, A.; Bhatt, S.; Shinde, U.; Tsivkovskii, R.; Blackburn, N.J.; Lutsenko, S. The lumenal loop Met672-Pro707 of copper-transporting ATPase ATP7A binds metals and facilitates copper release from the intramembrane sites. J. Biol. Chem. 2011, 286, 26585–26594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Köhn, B.; Shanmugavel, K.P.; Wu, M.; Kovermann, M.; Wittung-Stafshede, P. A Luminal Loop of Wilson Disease Protein Binds Copper and Is Required for Protein Activity. Biophys. J. 2018, 115, 1007–1018. [Google Scholar] [CrossRef] [PubMed]
  27. Vandecaetsbeek, I.; Vangheluwe, P.; Raeymaekers, L.; Wuytack, F.; Vanoevelen, J. The Ca2+ pumps of the endoplasmic reticulum and Golgi apparatus. Cold. Spring. Harb. Perspect. Biol. 2011, 3, a004184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Kline, C.D.; Gambill, B.F.; Mayfield, M.; Lutsenko, S.; Blackburn, N.J. pH-regulated metal-ligand switching in the HM loop of ATP7A: A new paradigm for metal transfer chemistry. Metallomics 2016, 8, 729–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Voskoboinik, I.; Mar, J.; Strausak, D.; Camakaris, J. The regulation of catalytic activity of the menkes copper-translocating P-type ATPase. Role of high affinity copper-binding sites. J. Biol. Chem. 2001, 276, 28620–28627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Petris, M.J.; Mercer, J.F.; Culvenor, J.G.; Lockhart, P.; Gleeson, P.A.; Camakaris, J. Ligand-regulated transport of the Menkes copper P-type ATPase efflux pump from the Golgi apparatus to the plasma membrane: A novel mechanism of regulated trafficking. EMBO J. 1996, 15, 6084–6095. [Google Scholar] [CrossRef] [PubMed]
  31. Hartwig, C.; Zlatic, S.A.; Wallin, M.; Vrail, A. Trafficking mechanisms of P-type ATPase copper transporters. Curr. Opin. Cell Biol. 2019, 59, 24–33. [Google Scholar] [CrossRef]
  32. Liu, Y.; Pilankatta, R.; Hatori, Y.; Lewis, D.; Inesi, G. Comparative features of copper ATPases ATP7A and ATP7B heterologously expressed in COS-1 cells. Biochemistry 2010, 49, 10006–10012. [Google Scholar] [CrossRef] [PubMed]
  33. Comstra, H.S.; McArthy, J.; Rudin-Rush, S.; Hartwig, C.; Gokhale, A.; Zlatic, S.A.; Blackburn, J.B.; Werner, E.; Petris, M.; D’Souza, P.; et al. The interactome of the copper transporter ATP7A belongs to a network of neurodevelopmental and neurodegeneration factors. eLife 2017, 6, e24722. [Google Scholar] [CrossRef] [PubMed]
  34. Roelofsen, H.; Wolters, H.; Van Luyn, M.J.; Miura, N.; Kuipers, F.; Vonk, R.J. Copper-induced apical trafficking of ATP7B in polarized hepatoma cells provides a mechanism for biliary copper excretion. Gastroenterology 2000, 119, 782–793. [Google Scholar] [CrossRef] [PubMed]
  35. Guo, Y.; Nyasae, L.; Braiterman, L.T.; Hubbard, A.L. NH2-terminal signals in ATP7B Cu-ATPase mediate its Cu-dependent anterograde traffic in polarized hepatic cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2005, 289, 12. [Google Scholar] [CrossRef] [Green Version]
  36. Fanni, D.; Pilloni, L.; Orrù, S.; Coni, P.; Liguori, C.; Serra, S.; Lai, M.L.; Uccheddu, A.; Contu, L.; Van Eyken, P.; et al. Expression of ATP7B in normal human liver. Eur. J. Histochem. 2005, 49, 371–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Petrukhin, K.; Lutsenko, S.; Chernov, I.; Ross, B.M.; Kaplan, J.H.; Gilliam, T.C. Characterization of the Wilson disease gene encoding a P-type copper transporting ATPase, genomic organization, alternative splicing, structure/function predictions. Hum. Mol. Genet. 1994, 3, 1647–1656. [Google Scholar] [CrossRef]
  38. Hatori, Y.; Inouye, S.; Akagi, R. Thiol-based copper handling by the copper chaperone Atox1. IUBMB Life 2017, 69, 246–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Yoo, S.K.; Cheong, J.; Kim, H.Y. STAMPing into Mitochondria. Int. J. Biol. Sci. 2014, 10, 321–326. [Google Scholar] [CrossRef] [Green Version]
  40. Sharp, P. The molecular basis of copper and iron interactions. Proc. Nutr. Soc. 2004, 63, 563–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Gulec, S.; Collins, J.F. Molecular mediators governing iron-copper interactions. Annu. Rev. Nutr. 2014, 34, 95–116. [Google Scholar] [CrossRef] [Green Version]
  42. Ohgami, R.S.; Campagna, D.R.; McDonald, A.; Fleming, M.D. The Steap proteins are metalloreductases. Blood 2006, 108, 1388–1394. [Google Scholar] [CrossRef]
  43. Vashchenko, G.; MacGillivray, R.T. Multi-copper oxidases and human iron metabolism. Nutrients 2013, 5, 2289–2313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Rouault, T.A. Biogenesis of iron-sulfur clusters in mammalian cells: New insights and relevance to human disease. Dis. Model. Mech. 2012, 5, 155–164. [Google Scholar] [CrossRef] [Green Version]
  45. Bhagi-Damodaran, A.; Michael, M.A.; Zhu, Q.; Reed, J.; Sandoval, B.A.; Mirts, E.N.; Chakraborty, S.; Moënne-Loccoz, P.; Zhang, Y.; Lu, Y. Why copper is preferred over iron for oxygen activation and reduction in haem-copper oxidases. Nat. Chem. 2017, 9, 257–263. [Google Scholar] [CrossRef]
  46. Collins, J.F.; Prohaska, J.R.; Knutson, M.D. Metabolic crossroads of iron and copper. Nutr. Rev. 2010, 68, 133–147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Nilsson, R.; Schultz, I.J.; Pierce, E.L.; Soltis, K.A.; Naranuntarat, A.; Ward, D.M.; Baughman, J.M.; Paradkar, P.N.; Kingsley, P.D.; Culotta, V.C.; et al. Discovery of genes essential for heme biosynthesis through large-scale gene expression analysis. Cell Metab. 2009, 10, 119–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Lill, R.; Mühlenhoff, U. Maturation of iron-sulfur proteins in eukaryotes: Mechanisms, connected processes, and diseases. Annu. Rev. Biochem. 2008, 77, 669–700. [Google Scholar] [CrossRef] [Green Version]
  49. Paul, B.T.; Manz, D.H.; Torti, F.M.; Torti, S.V. Mitochondria and Iron: Current questions. Expert. Rev. Hematol. 2017, 10, 65–79. [Google Scholar] [CrossRef] [Green Version]
  50. Xu, W.; Barrientos, T.; Andrews, N.C. Iron and Copper in Mitochondrial Diseases. Cell Metab. 2013, 17, 319–328. [Google Scholar] [CrossRef] [Green Version]
  51. Grunewald, T.G.; Bach, H.; Cossarizza, A.; Matsumoto, I. The STEAP protein family: Versatile oxidoreductases and targets for cancer immunotherapy with overlapping and distinct cellular functions. Biol. Cell. 2012, 104, 641–657. [Google Scholar] [CrossRef]
  52. Ohgami, R.S.; Campagna, D.R.; Greer, E.L.; Antiochos, B.; McDonald, A.; Chen, J.; Sharp, J.J.; Fujiwara, Y.; Barker, J.E.; Fleming, M.D. Identification of a ferrireductase required for efficient transferrin-dependent iron uptake in erythroid cells. Nat. Genet. 2005, 37, 1264–1269. [Google Scholar] [CrossRef] [Green Version]
  53. Gomes, I.M.; Maia, C.J.; Santos, C.R. STEAP proteins: From structure to applications in cancer therapy. Mol. Cancer Res. 2012, 10, 573–587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Gauss, G.H.; Kleven, M.D.; Sendamarai, A.K.; Fleming, M.D.; Lawrence, C.M. The crystal structure of six-transmembrane epithelial antigen of the prostate 4 (Steap4), a ferri/cuprireductase, suggests a novel interdomain flavin-binding site. J. Biol. Chem. 2013, 288, 20668–20682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Knutson, M.D. Steap proteins: Implications for iron and copper metabolism. Nutr. Rev. 2007, 65, 335–340. [Google Scholar] [CrossRef] [PubMed]
  56. Scarl, R.T.; Lawrence, C.M.; Gordon, H.M.; Nunemaker, C.S. STEAP4: Its emerging role in metabolism and homeostasis of cellular iron and copper. J. Endocrinol. 2017, 234, R123–R134. [Google Scholar] [CrossRef]
  57. Tanaka, Y.; Matsumoto, I.; Iwanami, K.; Inoue, A.; Minami, R.; Umeda, N.; Kanamori, A.; Ochiai, N.; Miyazawa, K.; Sugihara, M.; et al. Six-transmembrane epithelial antigen of prostate4 (STEAP4) is a tumor necrosis factor alpha-induced protein that regulates IL-6, IL-8, and cell proliferation in synovium from patients with rheumatoid arthritis. Mod. Rheumatol. 2012, 22, 128–136. [Google Scholar] [CrossRef] [PubMed]
  58. Xue, X.; Bredell, B.X.; Anderson, E.R.; Martin, A.; Mays, C.; Nagao-Kitamoto, H.; Huang, S.; Győrffy, B.; Greenson, J.K.; Hardiman, K.; et al. Quantitative proteomics identifies STEAP4 as a critical regulator of mitochondrial dysfunction linking inflammation and colon cancer. Proc. Natl. Acad. Sci. USA 2017, 114, E9608–E9617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ebe, H.; Matsumoto, I.; Osada, A.; Kurata, I.; Kawaguchi, H.; Kondo, Y.; Tsuboi, H.; Sumida, T. Splice variant of STEAP4 localizes in the nucleus, making it a possible transcriptional regulator of IL-6 production. Mod. Rheumatol. 2019, 29, 714–716. [Google Scholar] [CrossRef]
  60. Lane, D.J.; Bae, D.H.; Merlot, A.M.; Sahni, S.; Richardson, D.R. Duodenal cytochrome b (DCYTB) in iron metabolism: An update on function and regulation. Nutrients 2015, 7, 2274–2296. [Google Scholar] [CrossRef] [Green Version]
  61. Stoj, C.; Kosman, D.J. Cuprous oxidase activity of yeast Fet3p and human ceruloplasmin: Implication for function. FEBS Lett. 2003, 554, 422–426. [Google Scholar] [CrossRef] [Green Version]
  62. Rouault, T.A.; Zhang, D.L.; Jeong, S.Y. Brain iron homeostasis, the choroid plexus, and localization of iron transport proteins. Metab. Brain. Dis. 2009, 24, 673–684. [Google Scholar] [CrossRef] [Green Version]
  63. Jeong, S.Y.; David, S. Glycosylphosphatidylinositol-anchored ceruloplasmin is required for iron efflux from cells in the central nervous system. J. Biol. Chem. 2003, 278, 27144–27148. [Google Scholar] [CrossRef] [Green Version]
  64. McCarthy, R.C.; Kosman, D.J. Iron transport across the blood-brain barrier: Development, neurovascular regulation and cerebral amyloid angiopathy. Cell Mol. Life Sci. 2014, 72, 709–727. [Google Scholar] [CrossRef] [Green Version]
  65. Hellman, N.E.; Kono, S.; Mancini, G.M.; Hoogeboom, A.J.; de Jong, G.J.; Gitlin, J.D. Mechanisms of copper incorporation into human ceruloplasmin. J. Biol. Chem. 2002, 277, 46632–46638. [Google Scholar] [CrossRef] [Green Version]
  66. Sedlak, E.; Wittung-Stafshede, P. Discrete roles of copper ions in chemical unfolding of human ceruloplasmin. Biochemistry 2007, 46, 9638–9644. [Google Scholar] [CrossRef]
  67. Yang, X.L.; Miura, N.; Kawarada, Y.; Terada, K.; Petrukhin, K.; Gilliam, T.; Sugiyama, T. Two forms of Wilson disease protein produced by alternative splicing are localized in distinct cellular compartments. Biochem. J. 1997, 326, 897–902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Gitlin, J.D.; Schroeder, J.J.; Lee-Ambrose, L.M.; Cousins, R.J. Mechanisms of caeruloplasmin biosynthesis in normal and copper-deficient rats. Biochem. J. 1992, 282, 835–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Yoshida, K.; Furihata, K.; Takeda, S.; Nakamura, A.; Yamamoto, K.; Morita, H.; Hiyamuta, S.; Ikeda, S.; Shimizu, N.; Yanagisawa, N. A mutation in the ceruloplasmin gene is associated with systemic hemosiderosis in humans. Nat. Genet. 1995, 9, 267–272. [Google Scholar] [CrossRef]
  70. Kinebuchi, M.; Matsuura, A.; Kiyono, T.; Nomura, Y.; Kimura, S. Diagnostic copper imaging of Menkes disease by synchrotron radiation-generated X-ray fluorescence analysis. Sci. Rep. 2016, 6, 33247. [Google Scholar] [CrossRef] [Green Version]
  71. Matsuda, I.; Pearson, T.; Holtzman, N.A. Determination of apoceruloplasmin by radioimmunoassay in nutritional copper deficiency, Menkes’ kinky hair syndrome, Wilson’s disease, and umbilical cord blood. Pediatr. Res. 1974, 8, 821–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hahn, P.; Qian, Y.; Dentchev, T.; Chen, L.; Beard, J.; Harris, Z.L.; Dunaief, J.L. Disruption of ceruloplasmin and hephaestin in mice causes retinal iron overload and retinal degeneration with features of age-related macular degeneration. Proc. Nat. Acad. Sci. 2004, 101, 13850–13855. [Google Scholar] [CrossRef] [Green Version]
  73. Cherukuri, S.; Potla, R.; Sarkar, J.; Nurko, S.; Harris, Z.L.; Fox, P.L. Unexpected role of ceruloplasmin in intestinal iron absorption. Cell Metabol 2005, 2, 309–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Jiang, R.; Hua, C.; Wan, Y.; Jiang, B.; Hu, H.; Zheng, J.; Fuqua, B.K.; Dunaief, J.L.; Anderson, G.J.; David, S.; et al. Hephaestin and ceruloplasmin play distinct but interrelated roles in iron homeostasis in mouse brain. J. Nutr. 2015, 145, 1003–1009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hudson, D.M.; Curtis, S.B.; Smith, V.C.; Griffiths, T.A.; Wong, A.Y.; Scudamore, C.H.; Buchan, A.M.; MacGillivray, R.T. Human hephaestin expression is not limited to enterocytes of the gastrointestinal tract but is also found in the antrum, the enteric nervous system, and pancreatic {beta}-cells. Am. J. Physiol. Gastrointest. Liver Physiol. 2010, 298, G425–G432. [Google Scholar] [CrossRef] [PubMed]
  76. Wolkow, N.; Song, D.; Song, Y.; Chu, S.; Hadziahmetovic, M.; Lee, J.C.; Iacovelli, J.; Grieco, S.; Dunaief, J.L. Ferroxidase hephaestin’s cell-autonomous role in the retinal pigment epithelium. Am. J. Pathol. 2012, 180, 1614–1624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Jiang, B.; Liu, G.; Zheng, J.; Chen, M.; Maimaitiming, Z.; Chen, M.; Liu, S.; Jiang, R.; Fuqua, B.K.; Dunaief, J.L.; et al. Hephaestin and ceruloplasmin facilitate iron metabolism in the mouse kidney. Sci. Rep. 2016, 6, 39470. [Google Scholar] [CrossRef] [PubMed]
  78. Xu, X.; Pin, S.; Gathinji, M.; Fuchs, R.; Harris, Z.L. Aceruloplasminemia: An inherited neurodegenerative disease with impairment of iron homeostasis. Ann. N. Y. Acad. Sci. 2004, 1012, 299–305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Zheng, J.; Chen, M.; Liu, G.; Xu, E.; Chen, H. Ablation of hephaestin and ceruloplasmin results in iron accumulation in adipocytes and type 2 diabetes. FEBS Lett. 2018, 592, 394–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Zhao, L.; Hadziahmetovic, M.; Wang, C.; Xu, X.; Song, Y.; Jinnah, H.A.; Wodzinska, J.; Iacovelli, J.; Wolkow, N.; Krajacic, P.; et al. Cp/Heph mutant mice have iron-induced neurodegeneration diminished by deferiprone. J. Neurochem. 2015, 135, 958–974. [Google Scholar] [CrossRef] [PubMed]
  81. Nittis, T.; Gitlin, J.D. Role of copper in the proteosome-mediated degradation of the multicopper oxidase hephaestin. J. Biol. Chem. 2004, 279, 25696–25702. [Google Scholar] [CrossRef] [Green Version]
  82. Lee, S.M.; Attieh, Z.K.; Son, H.S.; Chen, H.; Bacouri-Haidar, M.; Vulpe, C.D. Iron repletion relocalizes hephaestin to a proximal basolateral compartment in polarized MDCK and Caco2 cells. Biochem. Biophys. Res. Commun. 2012, 421, 449–455. [Google Scholar] [CrossRef] [Green Version]
  83. Kuo, Y.M.; Su, T.; Chen, H.; Attieh, Z.; Syed, B.A.; McKie, A.T.; Anderson, G.J.; Gitschier, J.; Vulpe, C.D. Mislocalisation of hephaestin, a multicopper ferroxidase involved in basolateral intestinal iron transport, in the sex linked anaemia mouse. Gut 2004, 53, 201–206. [Google Scholar] [CrossRef] [Green Version]
  84. Chen, H.; Attieh, Z.K.; Syed, B.A.; Kuo, Y.M.; Stevens, V.; Fuqua, B.K.; Andersen, H.S.; Naylor, C.E.; Evans, R.W.; Gambling, L.; et al. Identification of zyklopen, a new member of the vertebrate multicopper ferroxidase family, and characterization in rodents and human cells. J. Nutr. 2010, 140, 1728–1735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Sharma, P.; Reichert, M.; Lu, Y.; Markello, T.C.; Adams, D.R.; Steinbach, P.J.; Fuqua, B.K.; Parisi, X.; Kaler, S.G.; Vulpe, C.D.; et al. Biallelic HEPHL1 variants impair ferroxidase activity and cause an abnormal hair phenotype. PLoS Genet. 2019, 15, e1008143. [Google Scholar] [CrossRef]
  86. Orlova, N.A.; Kovnir, S.V.; Vorobiev, I.I.; Gabibov, A.G.; Vorobiev, A.I. Blood Clotting Factor VIII: From Evolution to Therapy. Acta Naturae. 2013, 5, 19–39. [Google Scholar] [CrossRef]
  87. Mann, K.G.; Lawler, C.M.; Vehars, G.A.; Church, W.R. Coagulation Factor V contains copper ion. J. Biol. Chem. 1984, 259, 12949–12951. [Google Scholar] [CrossRef]
  88. Zheng, C.; Zhang, B. Combined deficiency of coagulation factors V and VIII: An update. Semin. Thromb. Hemost. 2013, 39, 613–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Appenzeller-Herzog, C.; Hauri, H.P. The ER-Golgi intermediate compartment (ERGIC): In search of its identity and function. J Cell Sci. 2006, 119, 2173–2183. [Google Scholar] [CrossRef] [Green Version]
  90. Zhang, B.; Cunningham, M.A.; Nichols, W.C.; Bernat, J.A.; Seligsohn, U.; Pipe, S.W.; McVey, J.H.; Schulte-Overberg, U.; de Bosch, N.B.; Ruiz-Saez, A.; et al. Bleeding due to disruption of a cargo-specific ER-to-Golgi transport complex. Nat. Genet. 2003, 34, 220–225. [Google Scholar] [CrossRef] [PubMed]
  91. Peng, C.-H.; Hsu, C.-H.; Wang, N.-L.; Jiang, C.-B. Spontaneous retroperitoneal hemorrhage n Menkes disease: A rare case report. Medicine 2018, 97, e9869. [Google Scholar] [CrossRef] [PubMed]
  92. Cobine, P.A.; Pierrel, F.; Bestwick, M.L.; Winge, D.R. Mitochondrial matrix copper complex used in metallation of cytochrome oxidase and superoxide dismutase. J. Biol. Chem. 2006, 281, 36552–36559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Cobine, P.A.; Moore, S.A.; Leary, S.C. Getting out what you put in: Copper in mitochondria and its impacts on human disease. Biochim. Biophys. Acta Mol. Cell. Res. 2021, 1868, 118867. [Google Scholar] [CrossRef] [PubMed]
  94. Horn, D.; Barrientos, A. Mitochondrial copper metabolism and delivery to cytochrome c oxidase. IUBMB Life 2008, 60, 421–429. [Google Scholar] [CrossRef] [Green Version]
  95. Jett, K.A.; Leary, S.C. Building the CuA site of cytochrome c oxidase: A complicated, redox-dependent process driven by a surprisingly large complement of accessory proteins. J. Biol. Chem. 2018, 293, 4644–4652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Timón-Gómez, A.; Nývltová, E.; Abriata, L.A.; Vila, A.J.; Hosler, J.; Barrientos, A. Mitochondrial cytochrome c oxidase biogenesis: Recent developments. Semin. Cell Dev. Biol. 2018, 76, 163–178. [Google Scholar] [CrossRef] [PubMed]
  97. Baker, Z.N.; Cobine, P.A.; Leary, S.C. The mitochondrion: A central architect of copper homeostasis. Metallomics 2017, 9, 1501–1512. [Google Scholar] [CrossRef] [PubMed]
  98. Bresgen, N.; Eckl, P.M. Oxidative stress and the homeodynamics of iron metabolis. Biomolecules 2015, 5, 808–847. [Google Scholar] [CrossRef]
  99. Bhattacharjee, A.; Yang, H.; Duffy, M.; Robinson, E.; Conrad-Antoville, A.; Lu, Y.W.; Capps, T.; Braiterman, L.; Wolfgang, M.; Murphy, M.P.; et al. The Activity of Menkes Disease Protein ATP7A Is Essential for Redox Balance in Mitochondria. J. Biol. Chem. 2016, 291, 16644–16658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Banci, L.; Bertini, I.; Cavallaro, G.; Ciofi-Baffoni, S. Seeking the determinants of the elusive functions of Sco proteins. FEBS J. 2011, 278, 2244–2262. [Google Scholar] [CrossRef]
  101. Leary, S.C.; Sasarman, F.; Nishimura, T.; Shoubridge, E.A. Human SCO2 is required for the synthesis of CO II and as a thiol-disulphide oxidoreductase for SCO1. Hum. Mol. Genet. 2009, 18, 2230–2240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Pacheu-Grau, D.; Wasilewski, M.; Oeljeklaus, S.; Gibhardt, C.S.; Aich, A.; Chudenkova, M.; Dennerlein, S.; Deckers, M.; Bogeski, I.; Warscheid, B.; et al. COA6 Facilitates Cytochrome c Oxidase Biogenesis as Thiol-reductase for Copper Metallochaperones in Mitochondria. J. Mol. Biol. 2020, 432, 2067–2079. [Google Scholar] [CrossRef] [PubMed]
  103. Morgello, S.; Peterson, H.D.; Kahn, L.J.; Laufer, H. Menkes kinky hair disease with ragged red fibers. Dev. Med. Child. Neurol. 1988, 30, 812–816. [Google Scholar] [CrossRef] [PubMed]
  104. Kodama, H.; Okabe, I.; Yanagisawa, M.; Kodama, Y. Copper deficiency in the mitochondria of cultured skin fibroblasts from patients with Menkes syndrome. J. Inherit. Metab. Dis. 1989, 124, 386–389. [Google Scholar] [CrossRef]
  105. Pedespan, J.M.; Jouaville, L.S.; Cances, C.; Letellier, T.; Malgat, M.; Guiraud, P.; Coquet, M.; Vernhet, I.; Lacombe, D.; Mazat, J.P. Menkes disease: Study of the mitochondrial respiratory chain in three cases. Eur. J. Paediatr. Neurol. 1999, 3, 167–170. [Google Scholar] [CrossRef]
  106. DiMauro, S.; Lombes, A.; Nakase, H.; Mita, S.; Fabrizi, G.M.; Tritschler, H.J.; Bonilla, E.; Miranda, A.F.; DeVivo, D.C.; Schon, E.A. Cytochrome c oxidase deficiency. Pediatr. Res. 1990, 28, 536–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Klinman, J.P.; Bonnot, F. Intrigues and Intricacies of the Biosynthetic Pathways for the Enzymatic Quinocofactors: PQQ, TTQ, CTQ, TPQ, and LTQ. Chem. Rev. 2013, 114, 4343–4365. [Google Scholar] [CrossRef] [Green Version]
  108. Finney, J.; Moon, H.J.; Ronnebaum, T.; Lantz, M.; Mure, M. Human copper-dependent amine oxidases. Arch. Biochem. Biophys. 2014, 546, 19–32. [Google Scholar] [CrossRef]
  109. Howard, M.; Farrar, C.A.; Sacks, S.H. Structural and functional diversity of collectins and ficolins and their relationship to disease. Semin. Immunopathol. 2018, 40, 75–85. [Google Scholar] [CrossRef]
  110. Vallet, S.D.; Guéroult, M.; Belloy, N.; Dauchez, M.; Ricard-Blum, S. A Three-Dimensional Model of Human Lysyl Oxidase, a Cross-Linking Enzyme. ACS Omega 2019, 4, 8495–8505. [Google Scholar] [CrossRef] [PubMed]
  111. Davidson, V.L. Protein-Derived Cofactors Revisited: Empowering Amino Acid Residues with New Functions. Biochemistry 2018, 57, 3115–3125. [Google Scholar] [CrossRef]
  112. Guillard, M.; Dimopoulou, A.; Fischer, B.; Morava, E.; Lefeber, D.J.; Kornak, U.; Wevers, R.A. Vacuolar H+-ATPase meets glycosylation in patients with cutis laxa. Biochim. Biophys. Acta. 2009, 1792, 903–914. [Google Scholar] [CrossRef]
  113. Khosrowabadi, E.; Kellokumpu, S. Golgi pH and Ion Homeostasis in Health and Disease. Rev. Physiol. Biochem. Pharmacol. 2020, 49. [Google Scholar] [CrossRef]
  114. Rodriguez-Pascual, F.; Rosell-Garcia, T. Lysyl Oxidases: Functions and Disorders. J. Glaucoma 2018, 27, S15–S19. [Google Scholar] [CrossRef]
  115. Brodsky, B.; Persikov, A.V. Molecular Structure of the Collagen Triple Helix. Adv. Protein Chem. 2005, 70, 301–339. [Google Scholar] [CrossRef]
  116. Palaniyar, N.; Zhang, L.; Kuzmenko, A.; Ikegami, M.; Wan, S.; Wu, H.; Korfhagen, T.R.; Whitsett, J.A.; McCormack, F.X. The role of pulmonary collectin n-terminal domains in surfactant structure, function, and homeostasis in vivo. J. Biol. Chem. 2002, 277, 26971–26979. [Google Scholar] [CrossRef] [Green Version]
  117. van de Wetering, J.K.; van Golde, L.M.; Batenburg, J.J. Collectins: Players of the innate immune system. Eur. J. Biochem. 2004, 271, 1229–1249. [Google Scholar] [CrossRef] [PubMed]
  118. Lintner, K.E.; Wu, Y.L.; Yang, Y.; Spencer, C.H.; Hauptmann, G.; Hebert, L.A.; Atkinson, J.P.; Yu, C.Y. Early Components of the Complement Classical Activation Pathway in Human Systemic Autoimmune Diseases. Front. Immunol. 2016, 7, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Casals, C.; García-Fojeda, B.; Minutti, C.M. Soluble defense collagens: Sweeping up immune threats. Mol. Immunol. 2019, 112, 291–304. [Google Scholar] [CrossRef] [PubMed]
  120. Lacroix, M.; Tessier, A.; Dumestre-Pérard, C.; Vadon-Le Goff, S.; Gout, E.; Bruckner-Tuderman, L.; Kiritsi, D.; Nyström, A.; Ricard-Blum, S.; Moali, C.; et al. Interaction of Complement Defence Collagens C1q and Mannose-Binding Lectin with BMP-1/Tolloid-like Proteinases. Sci. Rep. 2017, 7, 16958. [Google Scholar] [CrossRef] [PubMed]
  121. Khoriaty, R.; Vasievich, M.P.; Ginsburg, D. The COPII pathway and hematologic disease. Blood 2012, 120, 31–38. [Google Scholar] [CrossRef] [PubMed]
  122. Arakel, E.C.; Schwappach, B. Formation of COPI-coated vesicles at a glance. J. Cell Sci. 2018, 131, jcs209890. [Google Scholar] [CrossRef] [Green Version]
  123. Anelli, T.; Panina-Bordignon, P. How to Avoid a No-Deal ER Exit. Cells 2019, 8, 1051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Fu, Y.L.; Zhang, B.; Mu, T.W. LMAN1 (ERGIC-53) promotes trafficking of neuroreceptors. Biocem. Biophys. Acta 2019, 511, 356–362. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, B.; Stanford, K.R.; Kundu, M. ER-to-Golgi Trafficking and Its Implication in Neurological Diseases. Cells 2020, 9, 408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Donger, C.; Krejci, E.; Serradell, A.P.; Eymard, B.; Bon, S.; Nicole, S.; Chateau, D.; Gary, F.; Fardeau, M.; Massoulie, J.; et al. Mutation in the human acetylcholinesterase-associated collagen gene, COLQ, is responsible for congenital myasthenic syndrome with end-plate acetylcholinesterase deficiency (type Ic). Am. J. Hum. Genet. 1998, 63, 967–975. [Google Scholar] [CrossRef] [Green Version]
  127. Alders, M.; Hogan, B.M.; Gjini, E.; Salehi, F.; Al-Gazali, L.; Hennekam, E.A.; Holmberg, E.E.; Mannens, M.M.A.M.; Mulder, M.F.; Offerhaus, G.J.A.; et al. Mutations in CCBE1 cause generalized lymph vessel dysplasia in humans. Nat. Genet. 2009, 41, 1272–1274. [Google Scholar] [CrossRef]
  128. Yongqing, T.; Wilmann, P.G.; Reeve, S.B.; Coetzer, T.H.; Smith, A.I.; Whisstock, J.C.; Pike, R.N.; Wijeyewickrema, L.C. The x-ray crystal structure of mannose-binding lectin-associated serine proteinase-3 reveals the structural basis for enzyme inactivity associated with the Carnevale, Mingarelli, Malpuech, and Michels (3MC) syndrome. J. Biol. Chem. 2013, 288, 22399–22407. [Google Scholar] [CrossRef] [Green Version]
  129. Venkatraman Girija, U.; Furze, C.M.; Gingras, A.R.; Yoshizaki, T.; Ohtani, K.; Marshall, J.E.; Wallis, A.K.; Schwaeble, W.J.; El-Mezgueldi, M.; Mitchell, D.A.; et al. Molecular basis of sugar recognition by collectin-K1 and the effects of mutations associated with 3MC syndrome. BMC Biol. 2015, 13, 27. [Google Scholar] [CrossRef] [Green Version]
  130. Moon, H.J.; Finney, J.; Ronnebaum, T.; Mure, M. Human lysyl oxidase-like 2. Bioorg. Chem. 2014, 57, 231–241. [Google Scholar] [CrossRef]
  131. Trackman, P.C. Functional importance of lysyl oxidase family propeptide regions. J. Cell Commun. Signal. 2018, 12, 45–53. [Google Scholar] [CrossRef] [Green Version]
  132. Guo, D.C.; Regalado, E.S.; Gong, L.; Duan, X.; Santos-Cortez, R.L.; Arnaud, P.; Ren, Z.; Cai, B.; Hostetler, E.M.; Moran, R.; et al. LOX mutations predispose to thoracic aortic aneurysms and dissections. Circ. Res. 2016, 118, 928–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Lee, V.S.; Halabi, C.M.; Hoffman, E.P.; Carmichael, N.; Leshchiner, I.; Lian, C.G.; Bierhals, A.J.; Vuzman, D.; Brigham Genomic Medicine; Mecham, R.P.; et al. Loss of function mutation in LOX causes thoracic aortic aneurysm and dissection in humans. Proc. Nat. Acad. Sci. USA 2016, 113, 8759–8764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Zhang, X.; Wang, Q.; Wu, J.; Wang, J.; Shi, Y.; Liu, M. Crystal Structure of Human Lysyl Oxidase-like 2 (HLOXL2) in a Precursor State. Proc. Natl. Acad. Sci. USA 2018, 115, 3828–3833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Liu, X.; Zhao, Y.; Gao, J.; Pawlyk, B.; Starcher, B.; Spencer, J.A.; Yanagisawa, H.; Zuo, J.; Li, T. Elastic fiber homeostasis requires lysyl oxidase-like 1 protein. Nat. Genet. 2004, 36, 178–182. [Google Scholar] [CrossRef] [PubMed]
  136. Schlötzer-Schrehardt, U.; Zenkel, M. The role of lysyl oxidase-like 1 (LOXL1) in exfoliation syndrome and glaucoma. Exp. Eye Res. 2019, 189, 107818. [Google Scholar] [CrossRef]
  137. Pasutto, F.; Zenkel, M.; Hoja, U.; Berner, D.; Uebe, S.; Ferrazzi, F.; Schödel, J.; Liravi, P.; Ozaki, M.; Paoli, D.; et al. Pseudoexfoliation syndrome-associated genetic variants affect transcription factor binding and alternative splicing of LOXL1. Nat. Commun. 2017, 8, 15466. [Google Scholar] [CrossRef] [Green Version]
  138. Bignon, M.; Pichol-Thieved, C.; Hardouin, J.; Malbouyres, M.; Brechot, N.; Nasciutti, L.; Barret, A.; Teillon, J.; Guillon, E.; Etienne, E.; et al. Lysyl oxidase-like protein-2 regulates sprouting angiogenesis and type IV collagen assembly in the endothelial basement membrane. Blood 2011, 118, 3979–3989. [Google Scholar] [CrossRef] [PubMed]
  139. Abreu-Velez, A.M.; Howard, M.S. Collagen IV in Normal Skin and in Pathological Processes. N. Am. J. Med. Sci. 2012, 4, 1–8. [Google Scholar] [CrossRef] [Green Version]
  140. Herranz, N.; Dave, N.; Millanes-Romero, A.; Pascual-Reguant, L.; Morey, L.; Díaz, V.M.; Lórenz-Fonfría, V.; Gutierrez-Gallego, R.; Jerónimo, C.; Iturbide, A.; et al. Lysyl oxidase-like 2 (LOXL2) oxidizes trimethylated lysine 4 in histone H3. FEBS J. 2016, 283, 4263–4273. [Google Scholar] [CrossRef]
  141. Kashtan, C.E. Alport syndrome: An inherited disorder of renal, ocular, and cochlear basement membranes. Medicine 1999, 78, 338–360. [Google Scholar] [CrossRef]
  142. Cosgrove, D.; Dufek, B.; Meehan, D.T.; Delimont, D.; Hartnett, M.; Samuelson, G.; Gratton, M.A.; Phillips, G.; MacKenna, D.A.; Bain, G. Lysyl oxidase like-2 contributes to renal fibrosis in Col4α3/Alport mice. Kidney Internat. 2018, 94, 303–314. [Google Scholar] [CrossRef]
  143. Li, J.; Gao, B.; Xiao, X.; Li, S.; Jia, X.; Sun, W.; Guo, X.; Zhang, Q. Exome sequencing identified null mutations in LOXL3 associated with early-onset high myopia. Mol. Vis. 2016, 22, 161–167. [Google Scholar]
  144. Alzahrani, F.; Al Hazzaa, S.A.; Tayeb, H.; Alkuraya, F.S. LOXL3, encoding lysyl oxidase-like 3, is mutated in a family with autosomal recessive Stickler syndrome. Hum. Genet. 2015, 134, 451–453. [Google Scholar] [CrossRef]
  145. Ma, L.; Huang, C.; Wang, X.J.; Xin, D.E.; Wang, L.S.; Zou, Q.C.; Zhang, Y.S.; Tan, M.D.; Wang, Y.M.; Zhao, T.C.; et al. Lysyl Oxidase 3 Is a Dual-Specificity Enzyme Involved in STAT3 Deacetylation and Deacetylimination Modulation. Mol. Cell. 2017, 65, 296–309. [Google Scholar] [CrossRef] [Green Version]
  146. Peltonen, L.; Kuivaniemi, H.; Palotie, A.; Horn, N.; Kaitila, I.; Kivirikko, K.I. Alterations in copper and collagen metabolism in the Menkes syndrome and a new subtype of the Ehlers-Danlos syndrome. Biochemistry 1983, 22, 6156–6163. [Google Scholar] [CrossRef] [PubMed]
  147. Royce, P.M.; Steinmann, B. Markedly reduced activity of lysyl oxidase in skin and aorta from a patient with Menkes disease showing unusually severe connective tissue manifestations. Pediatr. Res. 1990, 28, 137–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Jankov, R.P.; Boerkoel, C.F.; Hellmann, J.; Sirkin, W.L.; Al-Maghrabi, J.; Tümer, Z.; Horn, N.; Feigenbaum, A. Lethal neonatal Menkes disease with severe vasculopathy and fractures. Acta Paediatr. 1998, 87, 1297–1300. [Google Scholar] [CrossRef]
  149. Bacopoulou, F.; Henderson, L.; Philip, S.G. Menkes disease mimicking non-accidental injury. Arch. Dis. Child. 2006, 91, 919. [Google Scholar] [CrossRef] [PubMed]
  150. Droms, R.J.; Rork, J.F.; McLean, R.; Martin, M.; Belazarian, L.; Wiss, K. Menkes Disease Mimicking Child Abuse. Pediatr. Dermatol. 2017, 34, e132–e134. [Google Scholar] [CrossRef]
  151. Bowie, D. Polyamine-mediated channel block of ionotropic glutamate receptors and its regulation by auxiliary proteins. J. Biol. Chem. 2018, 293, 18789–18802. [Google Scholar] [CrossRef] [Green Version]
  152. Nichols, C.G.; Lee, S.-J. Polyamines and potassium channels: A twenty five year romance. J. Biol. Chem. 2018, 293, 18779–18788. [Google Scholar] [CrossRef] [Green Version]
  153. Pegg, A.E. Functions of Polyamines in Mammals. J. Biol. Chem. 2016, 291, 14904–14912. [Google Scholar] [CrossRef] [Green Version]
  154. McGrath, A.P.; Hilmer, K.M.; Collyer, C.A.; Shepard, E.M.; Elmore, B.O.; Brown, D.E.; Dooley, D.M.; Guss, J.M. Structure and inhibition of human diamine oxidase. Biochemistry 2009, 48, 9810–9822. [Google Scholar] [CrossRef] [Green Version]
  155. Lieberman, P. Histamine, antihistamines, and the central nervous system. Allergy Asthma Proc. 2009, 30, 482–486. [Google Scholar] [CrossRef] [PubMed]
  156. Shahid, M.; Tripathi, T.; Sobia, F.; Moin, S.; Siddiqui, M.; Khan, B.A. Histamine, Histamine Receptors, and their Role in Immunomodulation: An Updated Systematic Review. Open Immunol. J. 2009, 2, 9–41. [Google Scholar] [CrossRef] [Green Version]
  157. Parsons, M.E.; Ganellin, C.R. Histamine and its receptors. Br. J. Pharmacol. 2006, 147, S127–S135. [Google Scholar] [CrossRef] [Green Version]
  158. Oosterwijk, J.C.; Richard, G.; van der Wielen, M.J.R.; van de Vosse, E.; Harth, W.; Sandkuijl, L.A.; Bakker, E.; van Ommen, G.-J.B. Molecular genetic analysis of two families with keratosis follicularis spinulosa decalvans: Refinement of gene localization and evidence for genetic heterogeneity. Hum. Genet. 1997, 100, 520–524. [Google Scholar] [CrossRef] [PubMed]
  159. Gimelli, G.; Giglio, S.; Zuffardi, O.; Alhonen, L.; Suppola, S.; Cusano, R.; Lo Nigro, C.; Gatti, R.; Ravazzolo, R.; Seri, M. Gene dosage of the spermidine/spermine N(1)-acetyltransferase (SSAT) gene with putrescine accumulation in a patient with a Xp21.1p22.12 duplication and keratosis follicularis spinulosa decalvans (KFSD). Hum. Genet. 2002, 111, 235–241. [Google Scholar] [CrossRef]
  160. Kaitaniemi, S.; Elovaara, H.; Groen, K.; Kidron, H.; Liukkonen, J.; Salminen, T.; Salmi, M.; Jalkanen, S.; Elima, K. The unique substrate specificity of human AOC2, a semicarbazide-sensitive amine oxidase. Cell. Mol. Life Sci. 2009, 66, 2743–2757. [Google Scholar] [CrossRef] [PubMed]
  161. Yang, H.; Ralle, M.; Wolfgang, M.J.; Dhawan, N.; Burkhead, J.L.; Rodriguez, S.; Kaplan, J.H.; Wong, G.W.; Haughey, N.; Lutsenko, S. Copper-dependent amino oxidase 3 governs selection of metabolic fuels in adipocytes. PLoS Biol. 2018, 16, e2006519. [Google Scholar] [CrossRef]
  162. Müller, J.W.; Gilligan, L.C.; Idkowiak, J.; Arlt, W.; Foster, P.A. The Regulation of Steroid Action by Sulfation and Desulfation. Endocr. Rev. 2015, 36, 526–563. [Google Scholar] [CrossRef]
  163. Schlotawa, L.; Adang, L.A.; Radhakrishnan, K.; Ahrens-Nicklas, R.C. Multiple Sulfatase Deficiency: A Disease Comprising Mucopolysaccharidosis, Sphingolipidosis, and More Caused by a Defect in Posttranslational Modification. Int. J. Mol. Sci. 2020, 21, 3448. [Google Scholar] [CrossRef] [PubMed]
  164. Schmidt, B.; Selmer, T.; Ingendoh, A.; von Figura, K. A novel amino acid modification in sulfatases that is defective in multiple sulfatase deficiency. Cell 1995, 82, 271–278. [Google Scholar] [CrossRef] [Green Version]
  165. Selmer, T.; Hallmann, A.; Schmidt, B.; Sumper, M.; von Figura, K. The evolutionary conservation of a novel protein modification, the conversion of cysteine to serinesemialdehyde in arylsulfatase from Volvox carteri. Eur. J. Biochem. 1996, 238, 341–345. [Google Scholar] [CrossRef]
  166. Appel, M.J.; Meier, K.K.; Lafrance-Vanasse, J.; Lim, H.; Tsai, C.L.; Hedman, B.; Hodgson, K.O.; Tainer, J.A.; Solomon, E.I.; and Bertozzi, C.R. Formylglycine-generating enzyme binds substrate directly at a mononuclear Cu(I) center to initiate O2 activation. Proc. Natl. Acad. Sci. USA 2019, 116, 5370–5375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Meury, M.; Knop, M.; Seebeck, F.P. Structural basis for copper-oxygen mediated C– H bond activation by the formylglycine-generating enzyme. Angew. Chem. Int. Ed. Engl. 2017, 56, 8115–8119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Roeser, D.; Preusser-Kunze, A.; Schmidt, B.; Gasow, K.; Wittmann, J.G.; Dierks, T.; von Figura, K.; Rudolph, M.G. A general binding mechanism for all human sulfatases by the formylglycine-generating enzyme. Proc. Natl. Acad. Sci. USA 2006, 103, 81–86. [Google Scholar] [CrossRef] [Green Version]
  169. Cosma, M.P.; Pepe, S.; Annunziata, I.; Newbold, R.F.; Grompe, M.; Parenti, G.; Ballabio, A. The multiple sulfatase deficiency gene encodes an essential and limiting factor for the activity of sulfatases. Cell 2003, 113, 445–456. [Google Scholar] [CrossRef]
  170. Dierks, T.; Schlotawa, L.; Frese, M.A.; Radhakrishnan, K.; von Figura, K.; Schmidt, B. Molecular basis of multiple sulfatase deficiency, mucolipidosis II/III and Niemann-Pick C1 disease—Lysosomal storage disorders caused by defects of non-lysosomal proteins. Biochim. Biophys. Acta 2009, 1793, 710–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Zito, E.; Fraldi, A.; Pepe, S.; Annunziata, I.; Kobinger, G.; Di Natale, P.; Ballabio, A.; Cosma, M.P. Sulfatase activities are regulated by the interaction of the sulfatase-modifying factor 1 with SUMF2. EMBO Rep. 2005, 6, 655–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Fraldi, A.; Zito, E.; Annunziata, F.; Lombardi, A.; Cozzolino, M.; Monti, M.; Spampanato, C.; Ballabio, A.; Pucci, P.; Sitia, R.; et al. Multistep, sequential control of the trafficking and function of the multiple sulfatase deficiency gene product, SUMF1 by PDI, ERGIC-53 and ERp44. Hum. Mol. Genet. 2008, 17, 2610–2621. [Google Scholar] [CrossRef] [PubMed]
  173. Diez-Roux, G.; Ballabio, A. Sulfatases and human disease. Annu. Rev. Genom. Hum. Genet. 2005, 6, 355–379. [Google Scholar] [CrossRef]
  174. Hung, Y.H.; Bush, A.I.; La Fontaine, S. Links between copper and cholesterol in Alzheimer’s disease. Front. Physiol. 2013, 4, 111. [Google Scholar] [CrossRef] [Green Version]
  175. Laverty, D.; Thomas, P.; Field, M.; Andersen, O.J.; Gold, M.G.; Biggin, P.C.; Gielen, M.; Smart, T.G. Crystal structures of a GABAA-receptor chimera reveal new endogenous neurosteroid-binding sites. Nat. Struct. Mol. Biol. 2017, 24, 977–985. [Google Scholar] [CrossRef]
  176. Miller, P.S.; Scott, S.; Masiulis, S.; De Colibus, L.; Pardon, E.; Steyaert, J.; Aricescu, A.R. Structural basis for GABAA receptor potentiation by neurosteroids. Nat. Struct. Mol. Biol. 2017, 24, 986–992. [Google Scholar] [CrossRef]
  177. Hung, Y.H.; Faux, N.G.; Killilea, D.W.; Yanjanin, N.; Firnkes, S.; Volitakis, I.; Ganio, G.; Walterfang, M.; Hastings, C.; Porter, F.D.; et al. Altered transition metal homeostasis in Niemann-Pick disease, type C1. Metallomics 2014, 6, 542–553. [Google Scholar] [CrossRef] [Green Version]
  178. Danks, D.M.; Campbell, P.E.; Stevens, B.J.; Mayne, V.; Cartwright, E. Menkes’s kinky hair syndrome. An inherited defect in copper absorption with widespread effects. Pediatrics 1972, 50, 188–201. [Google Scholar]
  179. Heyne, K.; Dörner, K.; Graucob, E.; Wiedemann, H.R. Monophyle Vakuolisierung von Promyelozyten bei Menkes-Syndrom (Trichopoliodystrophie). Klin. Padiatr. 1978, 190, 576–579. [Google Scholar]
  180. Sayın, S.; Ünal, S.; Çetin, M.; Gümrük, F. Vacuolization in Myeloid and Erythroid Precursors in a Child with Menkes Disease. Turk. J. Hematol. 2019, 36, 203–204. [Google Scholar] [CrossRef]
  181. Elias, P.M.; Williams, M.L.; Choi, E.H.; Feingold, K.R. Role of cholesterol sulfate in epidermal structure and function: Lessons from X-linked ichthyosis. Biochim. Biophys. Acta 2014, 1841, 353–361. [Google Scholar] [CrossRef] [Green Version]
  182. Galve, J.; Vicente, A.; González-Enseñat, M.A.; Pérez-Dueñas, B.; Cusí, V.; Møller, L.B.; Julià, M.; Domínguez, A.; Ferrando, J. Neonatal erythroderma as a first manifestation of Menkes disease. Pediatrics 2012, 130, e239–e242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Palmer, M.A.; Blakeborough, L.; Harries, M.; Haslam, I.S. Cholesterol homeostasis: Links to hair follicle biology and hair disorders. Exp. Dermatol. 2020, 29, 299–311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Foster, P.A.; Mueller, J.W. Insights into steroid sulfation and desulfation pathways. J. Mol. Endocrinol. 2018, 61, T273–T285. [Google Scholar] [CrossRef] [Green Version]
  185. Vendelboe, T.V.; Harris, P.; Zhao, Y.; Walter, T.S.; Harlos, K.; El Omari, K.; Christensen, H.E.M. The crystal structure of human dopamine-hydroxylase at 2.9 A resolution. Sci. Adv. 2016, 2, e1500980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Kumar, D.; Mains, R.E.; Eipper, B.A. From POMC and α-MSH to PAM, molecular oxygen, copper, and vitamin C. J. Mol. Endocrinol. 2015, 56, T63–T76. [Google Scholar] [CrossRef] [Green Version]
  187. Tang, Y.L.; Epstein, M.P.; Anderson, G.M.; Zabetian, C.P.; Cubells, J.F. Genotypic and haplotypic associations of the DBH gene with plasma dopamine beta-hydroxylase activity in African Americans. Eur. J. Hum. Genet. 2007, 15, 878–883. [Google Scholar] [CrossRef] [Green Version]
  188. Garland, E.M.; Biaggioni, I. Dopamine Beta-Hydroxylase Deficiency. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Eds.; Updated 25 April 2019; University of Washington: Seattle, WA, USA, 2003; pp. 1993–2021. [Google Scholar]
  189. Hussain, L.S.; Reddy, V.; Maani, C.V. Physiology, Noradrenergic Synapse; StatPearls Publishing: Treasure Island, FL, USA, 2020. [Google Scholar]
  190. Kim, C.H.; Leung, A.; Huh, Y.H.; Yang, E.; Kim, D.J.; Leblanc, P.; Ryu, H.; Kim, K.; Kim, D.W.; Garland, E.M.; et al. Norepinephrine deficiency is caused by combined abnormal mRNA processing and defective protein trafficking of dopamine beta-hydroxylase. J. Biol. Chem. 2011, 286, 9196–9204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Xiao, T.; Ackerman, C.M.; Carroll, E.C.; Jia, S.; Hoagland, A.; Chan, J.; Thai, B.; Liu, C.S.; Isacoff, E.Y.; Chang, C.J. Copper regulates rest-activity cycles through the locus coeruleus-norepinephrine system. Nat. Chem. Biol. 2018, 14, 655–663. [Google Scholar] [CrossRef]
  192. Qian, Y.; Tiffany-Castiglioni, E.; Harris, E.D. A Menkes P-type ATPase involved in copper homeostasis in the central nervous system of the rat. Brain Res. Mol. Brain Res. 1997, 48, 60–66. [Google Scholar] [CrossRef]
  193. Gerbasi, V.; Lutsenko, S.; Lewis, E.J. A mutation in the ATP7B copper transporter causes reduced dopamine beta-hydroxylase and norepinephrine in mouse adrenal. Neurochem. Res. 2003, 28, 867–873. [Google Scholar] [CrossRef]
  194. Christensen, N.J. The biochemical assessment of sympathoadrenal activity in man. Clin. Auton. Res. 1991, 1, 167–172. [Google Scholar] [CrossRef]
  195. Kaler, S.G.; Holmes, C.S. Catecholamine metabolites affected by the copper-dependent enzyme dopamine-beta-hydroxylase provide sensitive biomarkers for early diagnosis of menkes disease and viral-mediated ATP7A gene therapy. Adv. Pharmacol. 2013, 68, 223–233. [Google Scholar] [CrossRef] [PubMed]
  196. De, M.; Bell, J.; Blackburn, N.J.; Mains, R.E.; Eipper, B.A. Role for an essential tyrosine in peptide amidation. J. Biol. Chem. 2006, 281, 20873–20882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. El Meskini, R.; Culotta, V.C.; Mains, R.E.; Eipper, B.A. Supplying Copper to the Cuproenzyme Peptidylglycine α-Amidating Monooxygenase. J. Biol. Chem. 2003, 278, 12278–12284. [Google Scholar] [CrossRef] [Green Version]
  198. Bousquet-Moore, D.; Mains, R.E.; Eipper, B.A. Peptidylgycine α-amidating monooxygenase and copper: A gene-nutrient interaction critical to nervous system function. J. Neurosci. Res. 2010, 88, 2535–2545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Prohaska, J.R.; Gybina, A.A.; Broderius, M.; Brokate, B. Peptidylglycine-alpha-amidating monooxygenase activity and protein are lower in copper-deficient rats and suckling copper-deficient mice. Arch. Biochem. Biophys. 2005, 434, 212–220. [Google Scholar] [CrossRef] [Green Version]
  200. Xin, X.; Mains, R.E.; Eipper, B.A. Monooxygenase X, a Member of the Copper-dependent Monooxygenase Family Localized to the Endoplasmic Reticulum. J. Biol. Chem. 2004, 279, 48159–48167. [Google Scholar] [CrossRef] [Green Version]
  201. Costin, G.E.; Valencia, J.C.; Vieira, W.D.; Lamoreux, M.L.; Hearing, V.J. Tyrosinase processing and intracellular trafficking is disrupted in mouse primary melanocytes carrying the underwhite (uw) mutation. A model for oculocutaneous albinism (OCA) type 4. J. Cell Sci. 2003, 116, 3203–321213. [Google Scholar] [CrossRef] [Green Version]
  202. Wiriyasermkul, P.; Moriyama, S.; Nagamori, S. Membrane transport proteins in melanosomes: Regulation of ions for pigmentation. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183318. [Google Scholar] [CrossRef] [PubMed]
  203. Negroiu, G.; Branza-Nichita, N.; Costin, G.E.; Titu, H.; Petrescu, A.J.; Dwek, R.A.; Petrescu, S.M. Investigation of the intracellular transport of tyrosinase and tyrosinase related protein (TRP)-1. The effect of endoplasmic reticulum (ER)-glucosidases inhibition. Cell Mol. Biol. 1999, 45, 1001–1010. [Google Scholar]
  204. Matoba, Y.; Kumagai, T.; Yamamoto, A.; Yoshitsu, H.; Sugiyama, M. Crystallographic evidence that the dinuclear copper center of tyrosinase is flexible during catalysis. J. Biol. Chem. 2006, 281, 8981–8990. [Google Scholar] [CrossRef] [Green Version]
  205. Kanteev, M.; Goldfeder, M.; Chojnacki, M.; Adir, N.; Fishman, A. The mechanism of copper uptake by tyrosinase from Bacillus megaterium. J. Biol. Inorg. Chem. 2013, 18, 895–903. [Google Scholar] [CrossRef]
  206. Petris, M.J.; Stransak, D.; Mercer, J.F. The Menkes copper transporter is required for the activation of tyrosinase. Hum. Mol. Genet. 2000, 9, 2845–2851. [Google Scholar] [CrossRef] [Green Version]
  207. Branza-Nichita, N.; Petrescu, A.J.; Negroiu, G.; Dwek, R.A.; Petrescu, S.M. N-glycosylation processing and glycoprotein—Lessons from the tyrosinase-related proteins. Chem. Rev. 2000, 100, 4697–4712. [Google Scholar] [CrossRef]
  208. Halaban, R.; Cheng, E.; Svedine, S.; Aron, R.; Hebert, D.N. Proper folding and endoplasmic reticulum to golgi transport of tyrosinase are induced by its substrates, DOPA and tyrosine. J. Biol. Chem. 2001, 276, 11933–11938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Branza-Nichita, N.; Negroiu, G.; Petrescu, A.J.; Garman, E.F.; Platt, F.M.; Wormald, M.R.; Dwek, R.A.; Petrescu, S.M. Mutations at critical N-glycosylation sites reduce tyrosinase activity by altering folding and quality control. J. Biol. Chem. 2000, 275, 8169–8175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Setty, S.R.; Tenza, D.; Sviderskaya, E.V.; Bennett, D.C.; Raposo, G.; Marks, M.S. Cell-specific ATP7A transport sustains copper-dependent tyrosinase activity in melanosomes. Nature 2008, 454, 1142–1146. [Google Scholar] [CrossRef] [Green Version]
  211. Petrescu, S.M.; Branza-Nichita, N.; Negroiu, G.; Petrescu, A.J.; Dwek, R.A. Tyrosinase and glycoprotein folding: Roles of chaperones that recognize glycans. Biochemistry 2000, 39, 5229. [Google Scholar] [CrossRef] [PubMed]
  212. Bellono, N.W.; Escobar, I.E.; Lefkovith, A.J.; Marks, M.S.; Oancea, E. An intracellular anion channel critical for pigmentation. eLife 2014, 3, e04543. [Google Scholar] [CrossRef] [PubMed]
  213. Esposito, R.; D’Aniello, S.; Squarzoni, P.; Pezzotti, M.R.; Ristoratore, F.; Spagnuolo, A. New Insights into the Evolution of Metazoan Tyrosinase Gene Family. PLoS ONE 2012, 7, e35731. [Google Scholar] [CrossRef] [Green Version]
  214. Hirobe, T. Keratinocytes regulate the function of melanocytes. Dermatol. Sin. 2014, 32, 200–204. [Google Scholar] [CrossRef] [Green Version]
  215. Zecca, L.; Wilms, H.; Geick, S.; Claasen, J.H.; Brandenburg, L.O.; Holzknecht, C.; Panizza, M.L.; Zucca, F.A.; Deuschl, G.; Sievers, J.; et al. Human neuromelanin induces neuroinflammation and neurodegeneration in the rat substantia nigra: Implications for Parkinson’s disease. Acta Neuropathol. 2008, 116, 47–55. [Google Scholar] [CrossRef] [PubMed]
  216. Fedorow, H.; Tribl, F.; Halliday, G.; Gerlach, M.; Riederer, P.; Double, K.L. Neuromelanin in human dopamine neurons: Comparison with peripheral melanins and relevance to Parkinson’s disease. Prog. Neurobiol. 2005, 75, 109–124. [Google Scholar] [CrossRef]
  217. Pan, T.; Li, X.; Jankovic, J. The association between Parkinson’s disease and melanoma. Int. J. Cancer. 2011, 128, 2251–2260. [Google Scholar] [CrossRef] [PubMed]
  218. Roberts, D.S.; Linthicum, F.H. Distribution of melanocytes in the human cochlea. Otol. Neurotol. 2015, 36, e99–e100. [Google Scholar] [CrossRef] [Green Version]
  219. Gi, M.; Shim, D.B.; Wu, L.; Bok, J.; Song, M.H.; Choi, J.Y. Progressive hearing loss in vitamin A-deficient mice which may be protected by the activation of cochlear melanocyte. Sci. Rep. 2018, 8, 16415. [Google Scholar] [CrossRef]
  220. Khordadpoor-Deilamani, F.; Akbari, M.T.; Karimipoor, M.; Javadi, G. Sequence analysis of tyrosinase gene in ocular and oculocutaneous albinism patients: Introducing three novel mutations. Mol. Vis. 2015, 21, 730–735. [Google Scholar]
  221. Gasch, A.T.; Caruso, R.C.; Kaler, S.G.; Kaiser-Kupfer, M. Menkes’ syndrome: Ophthalmic findings. Ophthalmology 2002, 109, 1477–1483. [Google Scholar] [CrossRef]
  222. Babior, B.M.; Kipnes, R.S.; Curnutte, J.T. Biological defense mechanisms. The production by leukocytes of superoxide, a potential bactericidal agent. J. Clin. Investig. 1973, 52, 741–744. [Google Scholar] [CrossRef] [PubMed]
  223. Curtain, C.C.; Ali, F.; Volitakis, I.; Cherny, R.A.; Norton, R.S.; Beyreuther, K.; Barrow, C.J.; Masters, C.L.; Bush, A.I.; Barnham, K.J. Alzheimer’s disease amyloid-beta binds copper and zinc to generate an allosterically ordered membrane-penetrating structure containing superoxide dismutase-like subunits. J. Biol. Chem. 2001, 276, 20466–20473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Shibata, N.; Hirano, A.; Kobayashi, M.; Umahara, T.; Kawanami, T.; Asayama, K. Cerebellar superoxide dismutase expression in Menkes’ kinky hair disease: An immunohistochemical investigation. Acta Neuropathol. 1995, 90, 198–202. [Google Scholar] [CrossRef] [PubMed]
  225. Miao, L.; St Clair, D.K. Regulation of superoxide dismutase genes: Implications in disease. Free Radic. Biol. Med. 2009, 47, 344–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Crapo, J.D.; Oury, T.; Rabouille, C.; Slot, J.W.; Chang, L.-Y. Copper, zinc superoxide dismutase is primarily a cytosolic protein in human cells. Proc. Nat. Acad. Sci. USA 1992, 89, 10405–10409. [Google Scholar] [CrossRef] [Green Version]
  227. Petrovic, N.; Comi, A.; Ettinger, M.J. Copper incorporation into superoxide dismutase in Menkes lymphoblasts. J. Biol. Chem. 1996, 271, 28335–28340. [Google Scholar] [CrossRef] [Green Version]
  228. Walker, A.K.; Turner, B.J.; Atkin, J.D. Endoplasmic Reticulum Stress and Protein Misfolding in Amyotrophic Lateral Sclerosis. In Protein Misfolding Disorders: A Trip into the ER; Bentham Science Publishers: Oak Par, IL, USA, 2009; pp. 56–76. [Google Scholar] [CrossRef] [Green Version]
  229. Sala, F.A.; Wright, G.S.A.; Antonyuk, S.V.; Garratt, R.C.; Hasnain, S.S. Molecular recognition and maturation of SOD1 by its evolutionarily destabilised cognate chaperone hCCS. PLoS Biol. 2019, 17, e3000141. [Google Scholar] [CrossRef] [Green Version]
  230. Skopp, A.; Boyd, S.D.; Ullrich, M.S.; Liu, L.; Winkler, D.D. Copper–zinc superoxide dismutase (Sod1) activation terminates interaction between its copper chaperone (Ccs) and the cytosolic metal-binding domain of the copper importer Ctr1. Biometals 2019, 32, 695–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Fetherolf, M.M.; Boyd, S.D.; Taylor, A.B.; Kim, H.J.; Wohlschlegel, J.A.; Blackburn, N.J.; Hart, P.J.; Winge, D.R.; Winkler, D.D. Copper-zinc superoxide dismutase is activated through a sulfenic acid intermediate at a copper ion entry site. J. Biol. Chem. 2017, 292, 12025–12040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Boyd, S.D.; Liu, L.; Bulla, L.; Winkler, D.D. Quantifying the Interaction between Copper-Zinc Superoxide Dismutase (Sod1) and its Copper Chaperone (Ccs1). J. Proteom. Bioinform. 2018, 11, 473. [Google Scholar] [CrossRef] [PubMed]
  233. Kawamata, H.; Manfredi, G. Import, maturation, and function of SOD1 and its copper chaperone CCS in the mitochondrial intermembrane space. Antioxid. Redox Signal. 2010, 13, 1375–1384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Tafuri, F.; Ronchi, D.; Magri, F.; Comi, G.P.; Corti, S. SOD1 misplacing and mitochondrial dysfunction in amyotrophic lateral sclerosis pathogenesis. Front. Cell Neurosci. 2015, 9, 336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Kim, P.K.; Hettema, E.H. Multiple Pathways for Protein Transport to Peroxisomes. J. Mol. Biol. 2015, 427, 1176–1190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Kalel, V.C.; Erdmann, R. Unraveling of the Structure and Function of Peroxisomal Protein Import Machineries. In Proteomics of Peroxisomes, Subcellular Biochemistry; del Río, L., Schrader, M., Eds.; Springer: Singapore, 2018; p. 89. [Google Scholar]
  237. Islinger, M.; Voelkl, A.; Fahimi, H.D.; Schrader, M. The peroxisome: An update on mysteries 2.0. Histochem. Cell Biol. 2018, 150, 443–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Platta, H.W. The cycling peroxisomal targeting signal type 1—receptor Pex5p: Reaching the circle’s end with ubiquitin. Recept. Clin. Investig. 2014, 1, e69. [Google Scholar] [CrossRef]
  239. Islinger, M.; Li, K.W.; Seitz, J.; Völkl, A.; Lüers, G.H. Hitchhiking of Cu/Zn superoxide dismutase to peroxisomes—Evidence for a natural piggyback import mechanism in mammals. Traffic 2009, 10, 1711–1721. [Google Scholar] [CrossRef]
  240. Fransen, M.; Nordgren, M.; Wang, B.; Apanaset, O. Role of peroxisomes in ROS/RNS-metabolism: Implications for human disease. Biochim. Biophys. Acta Mol. Basis Dis. 2012, 1822, 1363–1373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Klouwer, F.C.; Berendse, K.; Ferdinandusse, S.; Wanders, R.J.; Engelen, M.; Poll-The, B.T. Zellweger spectrum disorders: Clinical overview and management approach. Orphanet. J. Rare Dis. 2015, 10, 151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Yokoyama, A.; Ohno, K.; Hirano, A.; Shintaku, M.; Kato, M.; Hayashi, K.; Kato, S. Cerebellar expression of copper chaperone for superoxide, cytosolic cu/zn-superoxide dismutase, 4-hydroxy-2-nonenal, acrolein and heat shock protein 32 in patients with menkes kinky hair disease: Immunohistochemical study. Yonago Acta Med. 2014, 57, 23–35. [Google Scholar]
  243. Horn, N.; Tønnesen, T.; Tümer, Z. Menkes Disease: An X-linked Neurological Disorder of the Copper Metabolism. Brain Pathol. 1992, 2, 351–362. [Google Scholar] [CrossRef] [PubMed]
  244. Hilton, J.B.; White, A.R.; Crouch, P.J. Metal-deficient SOD1 in amyotrophic lateral sclerosis. J. Mol. Med. 2015, 93, 481–487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Wright, G.S.; Antonyuk, S.V.; Hasnain, S.S. A faulty interaction between SOD1 and hCCS in neurodegenerative disease. Sci. Rep. 2016, 6, 27691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Karlsson, K.; Sandström, J.; Edlund, A.; Marklund, S.L. Turnover of extracellular-superoxide dismutase in tissues. Lab. Investig. 1994, 70, 705–710. [Google Scholar] [PubMed]
  247. Hu, Y.; Shah, P.; Clark, D.J.; Ao, M.; Zhang, H. Reanalysis of global proteomic and phosphoproteomic data identified a large number of glycopeptides. Anal Chem. 2018, 90, 8065–8071. [Google Scholar] [CrossRef] [PubMed]
  248. Itoh, S.; Ozumi, K.; Kim, H.W.; Nakagawa, O.; McKinney, R.D.; Folz, R.J.; Zelko, I.N.; Ushio-Fukai, M.; Fukai, T. Novel mechanism for regulation of extracellular SOD transcription and activity by copper: Role of antioxidant-1. Free Radic. Biol. Med. 2009, 46, 95–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Gottfredsen, R.H.; Larsen, U.G.; Enghild, J.J.; Petersen, S.V. Hydrogen peroxide induce modifications of human extracellular superoxide dismutase that results in enzyme inhibition. Redox Biol. 2013, 1, 24–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Folz, R.J.; Guan, J.; Seldin, M.F.; Oury, T.D.; Enghild, J.J.; Crapo, J.D. Mouse extracellular superoxide dismutase: Primary structure, tissue-specific gene expression, chromosomal localization and lung in situ hybridization. Am. J. Respir. Cell Mol. Biol. 1997, 17, 393–403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Gongora, M.C.; Lob, H.E.; Landmesser, U.; Guzik, T.J.; Martin, W.D.; Ozumi, K.; Wall, S.M.; Wilson, D.S.; Murthy, N.; Gravanis, M.; et al. Loss of Extracellular Superoxide Dismutase Leads to Acute Lung Damage in the Presence of Ambient Air: A Potential Mechanism Underlying Adult Respiratory Distress Syndrome. Am. J. Pathol. 2008, 173, 915–926. [Google Scholar] [CrossRef] [Green Version]
  252. Polshakov, V.I.; Mantsyzov, A.B.; Kozin, S.A.; Adzhubei, A.A.; Zhokhov, S.S.; van Beek, W.; Kulikova, A.A.; Indeykina, M.I.; Mitkevich, V.A.; Makarov, A.A. A Binuclear Zinc Interaction Fold Discovered in the Homodimer of Alzheimer’s Amyloid-beta Fragment with Taiwanese Mutation D7H. Chem. Int. Ed. Engl. 2017, 56, 11734–11739. [Google Scholar] [CrossRef] [PubMed]
  253. Multhaup, G.; Schlicksupp, A.; Hesse, L.; Beher, D.; Ruppert, T.; Masters, C.L.; Beyreuther, K. The amyloid precursor protein of Alzheimer’s disease in the reduction of copper(II) to copper(I). Science 1996, 271, 1406–1409. [Google Scholar] [CrossRef] [PubMed]
  254. Nakamura, M.; Shishido, N.; Nunomura, A.; Smith, M.A.; Perry, G.; Hayashi, Y.; Nakayama, K.; Hayashi, T. Three histidine residues of amyloid-beta peptide control the redox activity of copper and iron. Biochemistry 2007, 46, 12737–12743. [Google Scholar] [CrossRef]
  255. Nicolas, M.; Hassan, B.A. Amyloid precursor protein and neural development. Development 2014, 141, 2543–2548. [Google Scholar] [CrossRef] [Green Version]
  256. Caldwell, J.H.; Klevanski, M.; Saar, M.; Müller, U.C. Roles of the amyloid precursor protein family in the peripheral nervous system. Mech. Dev. 2013, 130, 433–446. [Google Scholar] [CrossRef]
  257. Dingwall, C. A copper-binding site in the cytoplasmic domain of BACE1 identifies a possible link to metal homoeostasis and oxidative stress in Alzheimer’s disease. Biochem. Soc. Trans. 2007, 35, 571–573. [Google Scholar] [CrossRef] [PubMed]
  258. Hussain, I.; Powell, D.J.; Howlett, D.R.; Chapman, G.A.; Gilmour, L.; Murdock, P.R.; Tew, D.G.; Meek, T.D.; Chapman, C.; Schneider, K.; et al. ASP1 (BACE2) cleaves the amyloid precursor protein at the beta-secretase site. Mol. Cell. Neurosci. 2000, 16, 609–619. [Google Scholar] [CrossRef] [PubMed]
  259. Munro, K.M.; Nash, A.; Pigoni, M.; Lichtenthaler, S.F.; Gunnersen, J.M. Functions of the Alzheimer’s Disease Protease BACE1 at the Synapse in the Central Nervous System. J. Mol. Neurosci. 2016, 60, 305–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Charlwood, J.; Dingwall, C.; Matico, R.; Hussain, I.; Johanson, K.; Moore, S.; Powell, D.J.; Skehel, J.M.; Ratcliffe, S.; Clarke, B.; et al. Characterization of the glycosylation profiles of Alzheimer’s beta-secretase protein Asp-2 expressed in a variety of cell lines. J. Biol. Chem. 2001, 276, 16739–16748. [Google Scholar] [CrossRef] [Green Version]
  261. Bieberich, E. Synthesis, Processing, and Function of N-glycans in N-glycoproteins. Adv. Neurobiol. 2014, 9, 47–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. Clinical Phenotypes of ATP7A-linked Disorders.
Table 1. Clinical Phenotypes of ATP7A-linked Disorders.
Clinical Type *AbreviationOMIMAge of OnsetDiagnostic PointerConnective Tissue Involvement Motor Function *Mental Function *Age of Death §
Menkes DiseaseMNK3094000–8 monthsKinky Hair; floppy infant; mimicry of severe metabolic disorders of lysosomes, mitochondria, and peroxisomes; dysautonomiaSevere; osteoporoses; hemorrhages; bladder and bowel diverticulaePoor mobility; no head controlSevere mental retardation<3 years
Long Surviving MenkesLS3094000–8 monthsHair changes; dysautonomia; floppy; initial symptoms similar to MNKSevere to moderatePoor mobility; limited head control Severe mental retardation<15 years
Moderate MenkesMOD3094003–12 monthsCoarse hair; dysautonomia; initial course milder; clinical diagnosis difficultPresent, but not obvious; skeletal dysplasia may occurWheelchair bound; cannot sit unsupportedMentally retarded<40 years
Mild MenkesMILD3094002–3 yearsCoarse hair; dysautonomia; initial course mild; clinical diagnosis difficultMinimal to moderate; skeletal dysplasia; Walk with difficulty and with use of aidModerately retarded; Slow and dull>40 years
Occipital Horn Syndrome #OHS304150ChildhoodCoarse hair; X-linked family history; connective tissue problems; muscle affection; dysautonomia,Exostoses on occipital bones; skeletal dysplasia; cutis laxa; hyperextensible joints; vascular complicationsWalk independently; appear clumsySlow or dull to normal IQ40–60 years
X-linked distal spinal muscular atrophy 3SMAX3300489AdulthoodX-linked family history of muscle wasting;
 
dysautonomia with prevalent adrenergic involvement
Minor connective tissue involvement; minor occipital horns may occurWalk independently; progressive distal motor neuropathyNormal IQ; normal fertility<80 years
* clinical types are poorly delineated with overlapping symptoms and functions within and between groups; # occipital horns depend on head control and can occur from two years; § connective tissue complications shorten lifespan; Cu treatment prolongs lifespan.
Table 2. Copper-dependent Enzymes and basic properties.
Table 2. Copper-dependent Enzymes and basic properties.
EnzymeOMIMEC-no.CofactorCu DonorCu Loading SiteCu ChaperoneSubcellular LocalizationATP7A-Linked Cu Deficiency Symptoms
ATOX1 regulated
Copper Pumps
ATP7A3000117.2.2.8Mg; ATP
Cu-ATOX1 allosteric
ATOX1
Cu-GSH
CytosolATOX1 piggy-backingSP, TGN, PMCu storage in tissues, low in brain, liver, plasma; S-Cu diagnostic after 1.5 mo
ATP7B6068827.2.2.8Mg; ATP
Cu-ATOX1 allosteric
ATOX1
Cu-GSH
CytosolATOX1 piggy-backingSP, TGN, secretory vesiclesLow activity in brain and liver; Fe accumulation; icterus, steatosis
Copper ReductasesSTEAP16044151.16.1.-Heme; NADCu-HisEC redoxNApPM,
endosomes
Cu and Fe accumulation on plasma membranes and vesicles; hypochromic anemia
STEAP26050941.16.1.-Heme; NADCu-HisEC redoxNApPM, Golgi
STEAP36096711.16.1.-Heme; NADPCu-HisEC redoxNApPM,
endosomes
STEAP4 611098 1.16.1.-Heme; NADPHCu-HisEC redoxNApPM, ER, Golgi
endosomes
nucleus, MIT
CYBDR 605745 1.-.-.-Heme; ascorbateCu-HisEC redoxNApPM
Copper OxidasesCP 117700 1.16.3.1Cu; ascorbateATP7BERGICNKECCP low in plasma; diagnostic after 1½ month; Cu and Fe storage
FV+VIII 612309 300841 1.16.3.1Cu; ascorbateATP7BERGICNKECMild clotting deficiency
HEPH 300167 1.16.3.1Cu; ascorbateATP7AERGICNKVesiclesCu and Fe intracellular storage; AMD
HEPH1L 618455 1.16.3.1Cu; ascorbateATP7AERGICNKcis-GolgiCutis laxa
COX
 
CuA (II)
 
 
CuB (I)
 
 
516040
 
 
516030
1.9.3.1 
 
Cu
 
 
Cu;Heme
 
Redox
SCO1;
SCO2;
COA6
COX11
 
 
IMS
 
 
IMS
 
 
COX17
 
 
COX17
IMMLow COX in brain and liver due to low Cu availability; High lactate in blood;
Leigh-like symptoms; COX defects; ragged red fibers; hypotonia
Copper Quinone Amine OxidasesLOX 153455 1.4.3.1Cu; LTQATP7Acis-GolgiHEPH1L redoxECNumerous connective tissue abnormalities: tortuous vessels, aortic aneurisms, and dissections, umbilical or inguinal hernias, bladder and bowel diverticulae, loose joint and skin, osteoporosis, lymphedema, lung infections; collectin defects with protein trafficking problems, and deficient activation of complement pathway; NAI-like; cataract
LOXL11534561.4.3.1Cu; LTQATP7Acis-GolgiRedox loading #EC
LOXL2 606663 1.4.3.1Cu; LTQATP7Acis-GolgiRedox loading #ER, nucleus *
LOXL3 607163 1.4.3.1Cu; LTQATP7Acis-GolgiRedox loading #ER, nucleus *
LOXL4 607318 1.4.3.1Cu; LTQATP7Acis-GolgiRedox loading #EC
AOC11046101.4.3.22Cu; TPQATP7Acis-GolgiRedox loading #ECIchthyosis, alopecia, inflammation, conjunctivitis, atopy, photophobia, keratitis, diarrhoea, gastrointestinal polyps
AOC26022681.4.3.21Cu; TPQATP7Acis-GolgiRedox loading #PM
AOC36037351.4.3.21Cu; TPQATP7Acis-GolgiRedox loading #PM
FGly
Generation
SUMF1 6079391.8.3.7Cu; CaATP7AERSUMF2SPGAG accumulation in tissues and urine; metachromasia, Alder Reilly anomaly; overlapping clinical features of multiple sulfatase deficiency (MSD) mimicking metachromatic leukodystrophy,
mucopolysaccharidosis (MPS), mucolipidosis (MLP), chondrodysplasia punctata, hydrocephalus
FGly
Activated Sulfatases
ARSA6075743.1.6.8FGly; CaNApNApNApLysosomes
ARSB6115423.1.6.12FGly; CaNApNApNApLysosomes
ARSD3000023.1.6.-FGly; CaNApNApNApLysosomes
ARSF3000033.1.6.-FGly; CaNApNApNApEC
ARSE3001803.1.6.-FGly; CaNApNApNApGolgi
ARSG6100083.1.6.1FGly; CaNApNApNApLysosomes
ARSH3005863.1.6.-FGly; CaNApNApNApPM
ARSI6100093.1.6.-FGly; CaNApNApNApEC
ARSJ6100103.1.6.-FGly; CaNApNApNApEC
ARSK6100113.1.6.-FGly; CaNApNApNApEC
GALNS6122223.1.6.4FGly; CaNApNApNApLysosomes
GNS6076643.1.6.14FGly; CaNApNApNApLysosomes
IDS3008233.1.6.13FGly; CaNApNApNApLysosomes
STS3007473.1.6.2FGly; CaNApNApNApERichthyosis, seborrhoea, hair changes
Copper Amine OxidasesDBH6093121.14.17.1Cu; ascorbateATP7AERMOXD1#Vesicles; ECHigh DA/NE; vomiting, hypotension, hypothermia, hypoglycemia
PAM
PHM
PAL
170270 
1.14.17.3
4.3.2.5
Cu; ascorbate
Zn; Ca
 
ATP7A
NAp
 
TGN
NAp
 
NK
NAp
Golgi VesiclesPain, seizure, anxiety, impaired wakening, temperature, weight and fluid balance
TYR6069331.14.18.1Cu; ascorbateATP7AERTYRP1
TYRP2
MelanosomesAlbinism, visual and hearing problems
MOXD16090001.14.17.-Cu; ascorbateATP7AER-ER~DBH deficiency
Cu/Zn Superoxide DismutasesSOD11474501.15.1.1Cu; Zn
Cu-CCS
allosteric
GSH
ATP7A
matrix
NK
Cytosol
ER
IMS
NK
CCS redox and piggy-backing
CCS redox
CCS
Cytosol
peroxisomes
IMS
nucleus
Low SOD1 activity in nerve tissue and liver due to poor Cu availability; peroxisomal pathologies; motor neuron disease
SOD31854901.15.1.1Cu; ZnATP7ASPNKECLung disease, angiopathy
CCS603864NACu;ZnGSH
ATP7A
matrix
nucleus
Cytosol
ER
IMS
nucleus
NApCytosol
peroxisomes
IMS
nucleus
Purkinje cell pathologies; ALS-like phenotype
APP104760NACu; ZnNKSPNKECSenecense, cerebral angiopathy
APLP1104775NACu; ZnNKSPNKEC
APLP2104776NACu; ZnNKSPNKEC
Cu-CCS
Regulated Enzyme
BACE16042523.4.23.46Cu-CCS
allosteric
CCSSPCCSTGNPoor neuronal growth
#: indicated; *: histone biology; EC: extracellular; ER: endoplasmic reticulum; ERGIC: ER–Golgi intermediate compartment; GAG: glycoamino glycans; GHS: glutathione; IMM: inner mitochondrial membrane; IMS: intra mitochondrial space; MIT: mitochondria; NA: not assigned; NAI: non-accidental injuries; NAp: not applicable; NK: not known; PM: plasma membrane; SP: secretory pathway; TGN: trans-Golgi network.
Table 3. Menkes Disease Symptoms.
Table 3. Menkes Disease Symptoms.
Birth and Neonatal PeriodEnzyme DeficiencyComments
PretermAOC (histaminase)One-third before 37 weeks
Premature rupture of fetal membranes LOX
Weight-One-third less than 2500 g 
Length-
Head circumference-
Apgar score- Quick test at 1 and 5 min, in rare cases, also 10 min after birth
Denver scale-Developmental score for milestones in young children according to age
Bayley score- Cognitive, language, and motor developmental infants and toddlers score
Hydrops fetalis LOX Severe swelling (oedema)
Intrauterine growth retardationLOX, SUMF1Small for gestational age
Decreased fetal movements-
Neonatal onset-Rarely recognized before hair changes at 2–3 month
Neonatal death-
Early death - Usually before three years
Failure to thrive-
Feeding difficultiesDBH, LOXPoor sucking and swallowing
Floppy infantCOX, LOX
Poor head controlCOX
DysautonomyDBH
Infantile spasmsCOXShivers or a small jerks in series
Irritability DBH, PAM, SUMF1
Babinski reflex DBH, PAM, SUMF1Upward movement of the big toe sign of pyramidal dysfunction
AnxietyDBH, PAM, SUMF1
Increased response to noiseDBH, PAM, SUMF1
Lethargy DBH, PAM, SUMF1Decreased alertness
Respiratory distress LOX, SOD3
Icterus/jaundiceCP, LOXPhoto therapy resistant
External features
- Head and neck
Face lacking in expressionDBHLow mimic
PallorTYRLight skin color
HypertelorismLOX, SUMF1Widely spaced eyes
NystagmusDBH, LOX, TYRDifficulty in controlling eye movements
BlepharophimosisLOXNarrowing of eye opening
PhotophobiaAOC, TYRLight intolerance
KeratitisAOC, LOXCornea inflammation
ConjunctivitisAOC, LOXEye inflammation
PtosisDBH, LOXDroopy eyelids
MiosisDBH Excessive constriction of pupils
High arched (cupid) eyebrowsLOX, SUMF1
OphthalmoplegiaDBH, LOX
Cherubic appearanceLOX, SUMF1
Microcephaly LOX, SUMF1<2 SD for age
Brachycephaly LOX, SUMF1
Frontal bossingLOX, SUMF1
Occipital bossing LOX, SUMF1
Long philtrumLOX, SUMF1
High foreheadLOX, SUMF1
High-arched palateLOX, SUMF1
Small chinLOX, SUMF1
Pudgy cheeks SUMF1
Flat central faceLOX, SUMF1
Depressed nasal bridgeLOX, SUMF1
Nasal congestionLOX
Hypoplastic mandiblesLOX, SUMF1
MicrognathiaLOX, SUMF1
Retrognathia LOX, SUMF1
Drooping jawsSUMF1
Low set earsLOX, SUMF1
Large earsLOX, SUMF1
Occipital exostosesLOXCalcified exostoses palpable from occiput, uncommon
Internal jugular phlebectasiaLOX
- Chest
Pectus excavatumLOX
Pectus carinatumLOX
Neurological symptoms
Corpus callosum agenesis SUMF1 Absence of brain structure that connects the two hemispheres
DysautonomiaDBH
Cerebellar hypoplasiaLOX
Mental retardationCOX, PAM, SOD1
Motor retardationCOX, PAM, SOD1
Loss of milestones-Progressive neurologic defects
HypothermiaDBH, PAM Subnormal body temperature
HypoglycemiaDBH, PAM Subnormal sugar values
Nasal congestionDBH
West syndrome COX Epileptic encephalopathy
Seizures COX Refractory and early onset
Clonic seizures COX
Myoclonic seizures COX
Tonic seizures COX
Motor dysfunctionDBH
AtaxiaDBH, PAM, SOD1, SUMF1
SpasticityCOX
Hypertonia DBH
HypotoniaDBH
Eye symptoms
CataractLOX
MyopiaLOX
NystagmusDBH, LOX, TYRDifficulty in controlling eye movements
StrabismusTYR
BlepharophimosisLOXNarrowing of eye opening
PhotophobiaAOC, TYRLight intolerance
KeratitisAOC, LOXCornea inflammation
ConjunctivitisAOC, LOX
PtosisDBH, LOXDroopy eyelids
MiosisDBH Excessive constriction of pupils
Reduced visual acuityTYR
Optic discs palorTYR
Optic atrophy TYR Abnormal electroretinogram (ERG)
Visual loss TYR Visual evoked potential (VIP)
Retinal and iris depigmentationTYR
Iris trans-luminescenceTYR
Iris microcysts SUMF1, TYR
Hypopigmented fundus TYR Fundoscopy
Ear symptoms
Hearing loss LOX, PAM, TYR Brain stem auditory evoked potential (BAEP)
Hair and skin symptoms
Fine, silvery and brittle hair AOC, TYR, SUMF1Short, stubby, friable
Depigmented scalp hairTYR, SUMF1Lusterless, silvery, steel wool
Sparse hairAOC, SUMF1Rubbing against pillow may feel like unshaven stubbles
AlopeciaAOC, SUMF1Lack of hair
Fetal hair may be unaffected-Soft
Pili tortiSUMF1Hair twisted about their own axis
Trichorhexis nodosaSUMF1Frying and splitting of hair ends
Monilethrix SUMF1Varying diameters of the shafts
Cupid eyebrowsLOX, SUMF1Eyebrows with a high arch
Sparse eyebrowsAOC, SUMF1Look like old man’s eyebrows
Sparse eyelashesAOC, SUMF1Breaks easily
SeborrheaAOC, SUMF1Dry and scaly skin
Erythroderma AOC, SUMF1 Generalized exfoliative dermatitis with redness and scaling
Cutis laxaHEPHL1, LOXLax and wrinkled skin may give a progeria like appearance
Pale skinPAM, TYRAlmost like an albino
AnhydrosisDBH, LOX Inability to sweat normally
Doughy skinLOX Swelling of subcutaneous tissue
LymphedemaLOX Swelling due to poor lymphatic system
Dentation
Hyperplastic gumsLOXProminent gums
Dental abnormalitiesLOX
Enamel defectsLOX
Delayed eruptionLOX
Biconically shaped incisorsLOX
Lung symptoms
Acute respiratory distress syndrome AOC, LOX, SOD3, SUMF1
Chronic obstructive pulmonary disease AOC, LOX, SOD3, SUMF1
Emphysema LOX, SOD3, SUMF1 Damaged air sacs (alveoli) with breathing difficulty
Cardiovascular symptoms
Congenital heart disease COX, LOX About 5%
Angiopathy AOC, APP, LOX, SOD3 Disease of arteries, veins, and capillaries
Tortuous blood vessels LOX Twisted with frayed and split inner walls
Bleeding tendencyFV+VIII, LOX
Mild coagulation deficiencyFV+VIII
HematomasLOX
Subdural hematomas LOX
Intracranial hemorrhage LOX
CephalohematomasLOXPrevalent at birth
Gastrointestinal symptoms
Chronic diarrheaAOC
VomitingAOC
Bowel dysfunctionAOC
Gastrointestinal polypsLOX
Hiatal herniaLOX
Hepatic symptoms
HepatomegalyCOX, SOD1, SUMF1Low hepatic copper gives low enzymatic activity
IcterusCP, ATP7B Yellowish color of skin and eyes
SteatosisCOX, SOD1Fatty liver
Genitourinary symptoms
Bladder diverticulaLOX
Bladder ruptureLOX
Ureteral obstructionLOX
GlomerulonephritisLOX
Urinary tract infection LOX
Vesico-ureteral reflux LOX
HydronephrosisLOX Partial urinary tract blockage
Diaphragmatic hernia LOX
Umbilical herniasLOX
Inguinal hernia LOX
Cryptorchidism LOX, SUMF1 Undescended t esticles
Connective tissue symptoms
Loose/hypermobile jointsLOX
Tortuous vesselsLOX
Wrinkled and loose/extensible skinLOX
Soft skin / edemaLOX
Musculoskeletal symptoms
- Skeletal—neck and chest
Cervical spine anomalies LOXMimics non-accidental lesions
Short, broad claviclesLOX
Flaring of the ribsLOX
Short, broad ribsLOX
Pectus excavatumLOX Sunken breastbone
Pectus carinatumLOX Protruding breastbone; “pigeon chest”
- Skeletal—limbs
Congenital bone fractures LOXSymmetrical uncommon in “battered child”/NAI
Long-bone fracturesLOX
Metaphyseal spurring LOXCan resemble scurvy
Diaphyseal periosteal reactionLOX
Cortical thickening LOX
Short humeriLOX
- Skeletal—others
Wormian bonesLOXIntrasutural supernumerary bones, not found in child abuse
Spondylolysis LOX, SUMF1Fractures of vertebra
OsteroporosisLOX, SUMF1Brittle bones
OsteopeniaLOX, SUMF1
Cartilage malformationLOX, SUMF1
Joint laxityLOX
Limb dislocationsLOX
Metaphyseal widening LOX
OsteochondrodysplasiaLOX, SUMF1
Occipital horn exostosesLOXUncommon in MNK, but can be observed from 2 years
- Muscles
Motor neuron disease SOD1, LOX
InvestigationsActivity measured
MR MRI and MRA Neuroimaging Magnetic Resonance and computerized tomography; cerebral atrophy, cortical areas of low density, diffuse cerebral and cerebellar volume loss, white-matter, and basal ganglia changes
CT Neuroimaging
EEG Brain activity Hypsarrhythmia, diffuse, multifocal spike activity
Radiography Bone Symmetrical metaphyseal flaring and spurring of ribs, and cervical fractures may mimic non-accidental trauma, but these are not symmetrical; skull wormian bones are not seen in child abuse
ArteriographyVasculature Elongated and tortuous cerebral and systemic vessels
Ultrasonography Bladder, bowelDiverticulae and polyps
Light MicroscopyHair examinationPili torti, trichorexis nodosa, monilethrix
Echocardiography HeartHeart murmur
ERGElectroretinogram Optic atrophy
VIP Visual evoked potential Loss of vision; retinal and macular degeneration
Fundoscopy Eye background, macula Hypopigmented
BAEP Brain stem auditory evoked potential Hearing loss
Cell culture Radioactive copper testIncreased accumulation and retention
Tissue copperICPMS; AAIncreased in CVS, placenta, muscle; liver low
Biomarkers
BoyATP7AX-linked
Family history of male infant deathATP7AX-linked
HyperbilirubinemiaATP7BTransient, but prolonged and light therapy resistant
Low plasma copperATP7ADiagnostic from 4–6 weeks
Low free CuATP7ADiagnostic from birth
Low ceruloplasminCPDiagnostic from 4–6 weeks
High RBC (Red Blood Cells) CuSOD1Erythrocyte SOD1 in neonates
AnemiaHEPH, CPMay be hypochrome
NeutropeniaLOXDecreased neutrophils
ThromboembolismFV+VIIIBlood clot breaking loose and plugs other vessels
Urinary Cu low to normalATP7AMT
Low liver CuATP7ADiagnostic from birth
High placenta CuATP7ADiagnostic from birth; CVS diagnostic prenatally
High metallothionein levelsATP7AMT1 and MT2 (diagnostic?)
Plasma DA/NE ratio increased DBHDiagnostic from birth
Urinary HVA/VMA ratio increased DBHDiagnostic from birth
HypoglycemiaSUMF1, PAM, DBHTransient
High blood lactateCOXCSF, intermittent
PyruvateCOX Intermittent
HyperammoniemiaCOXIntermittent
High plasma glutamic acidCOXIntermittent, alpha-ketogluterate conversion
Respiratory chain deficienciesCOXIndicative
Intracellular Cu accumulationATP7ADiagnostic, tissue culture
Molecular screening of ATP7AATP7ADefinitive diagnosis
Pathology
Purkinje cell pathologies SOD1Faulty arborization and “weeping willow”
Ragged red fibersCOXSubsarcolemmal aggregates of mitochondria in muscle fiber
Alder Reilly anomaly SUMF1Vacuolization of blood cells; observed in GAG deficiencies
MetachromasiaSUMF1Color staining change of accumulated tissue sugar sulfatides
Pili tortiSUMF1Hair twisted about their own axis
Trichorhexis nodosaSUMF1Frying and splitting of hair ends
Monilethrix SUMF1Varying diameters of the shafts
Differential diagnosis:
NAILOXNon-accidental injuries; >10% symmetric changes think MNK
Osteogenesis imperfectaLOXBrittle bones and bone dysplasias
Mitochondrial disorder COXCompromised energy production affecting all organs and tissues
Organic acid uriaCOXDefective mitochondrial matrix metabolism
Cutis laxaLOXLoose and wrinkled skin in an infant
Progeria LOX, SUMF1Old age syndrome in young people
Syndromes with hair abnormalitiesSUMF1, AOC
Glutamine defectsCOXDefective mitochondrial matrix metabolism
MSDSUMF1Multiple sulfatase deficiency
MPSSUMF1Mucopolysaccaridoses
MLPSUMF1Mucolipidoses
LeukodystrophySUMF1E.g., metachromatic leukodystrophy
DBH deficiency, congenital DBHCNS Cu deficiency
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Horn, N.; Wittung-Stafshede, P. ATP7A-Regulated Enzyme Metalation and Trafficking in the Menkes Disease Puzzle. Biomedicines 2021, 9, 391. https://doi.org/10.3390/biomedicines9040391

AMA Style

Horn N, Wittung-Stafshede P. ATP7A-Regulated Enzyme Metalation and Trafficking in the Menkes Disease Puzzle. Biomedicines. 2021; 9(4):391. https://doi.org/10.3390/biomedicines9040391

Chicago/Turabian Style

Horn, Nina, and Pernilla Wittung-Stafshede. 2021. "ATP7A-Regulated Enzyme Metalation and Trafficking in the Menkes Disease Puzzle" Biomedicines 9, no. 4: 391. https://doi.org/10.3390/biomedicines9040391

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