Many viruses require the host endoplasmic reticulum protein-folding machinery in order to correctly fold one or more of their glycoproteins. Iminosugars with glucose stereochemistry target the glucosidases which are key for entry into the glycoprotein folding cycle. Viral glycoproteins are thus prevented from interacting with the protein-folding machinery leading to misfolding and an antiviral effect against a wide range of different viral families. As iminosugars target host enzymes, they should be refractory to mutations in the virus. Iminosugars therefore have great potential for development as broad-spectrum antiviral therapeutics. We outline the mechanism giving rise to the antiviral activity of iminosugars, the current progress in the development of iminosugar antivirals and future prospects for this field.

Enveloped viruses have a membrane surrounding their capsid, which is composed of lipids derived from host membranes and viral glycoproteins. These viral glycoproteins are typically dependent upon the host endoplasmic reticulum (ER) protein-folding machinery to form the three-dimensional structure necessary for secretion and/or activity [13]. Partial inhibition of the ER-folding mechanisms in order to prevent properly folded viral glycoproteins from being incorporated into the virus is therefore a promising target for the development of broad-spectrum antivirals [4,5].

That effective treatments for viral diseases are an urgent requirement has been highlighted by the recent outbreaks of Ebola and Zika. Whilst vaccines are considered the gold standard in this area, the production of effective vaccines can in some cases take many years and factors such as the presence of multiple viral serotypes, as seen for dengue virus (DENV), can pose an added challenge for vaccine development [6]. A complementary approach to vaccination is the development of antiviral drugs. Antiviral drugs fall into two classes: direct-acting drugs, which target one of the essential components of the virus itself, and host-directed drugs, which target a host process required for viral replication. The development of host-directed antivirals is challenging due to the need to ensure selective toxicity towards the virus; however, they avoid the problem of resistance mutations arising in the viral target, which can limit the efficacy of direct-acting antivirals (see, for example, recent reviews [7,8]).

The iminosugar class of compounds includes promising candidates for the development of host-directed antivirals (Figure 1). Iminosugars are sugar mimetics in which the cyclic oxygen is replaced with a nitrogen [9]. Their structural similarity to sugar molecules means that many iminosugars are competitive inhibitors of enzymes that act on sugar substrates. As such, their therapeutic potential has been explored against a range of diseases including Gaucher's disease for which the enzymatic target is glucosylceramide synthase [10,11], and type II diabetes where the targets are intestinal α-glucosidases [12]. Key host enzyme targets for antiviral activity are the ER α-glucosidases I and II (α-glu I and α-glu II), which control entry to the calnexin cycle, which is required for the folding of many glycoproteins [2,1317]. Iminosugars that inhibit these enzymes have antiviral activity against a range of enveloped viruses including DENV [18,19], influenza virus [20,21], hepatitis C virus (HCV) [22,23] and human immunodeficiency virus [24,25]. Support for the approach of targeting α-glu I in antiviral development is provided by the observations that two patients deficient in this enzyme have no clinical evidence of recurrent viral infections and that cells derived from these patients have a greatly reduced ability to support replication of a variety of different viruses [26]. The α-glu I-deficient patients present with multiple neurologic complications; however, such effects are not expected from antiviral therapy targeting α-glu I as only partial inhibition of the enzyme is necessary for an antiviral effect. In addition, antiviral treatment will not be administered during early development where a lack of α-glu I may be more likely to influence neurologic parameters. The sensitivity of viruses to inhibition of α-glu I and II appears to exceed that of the host proteins, giving a potential therapeutic window in which an antiviral effect occurs without a detrimental effect on the host.

It is widely accepted that a key mechanism by which iminosugars act as antivirals is their ability to disrupt glycoprotein folding, arising from the inhibition of ER α-glu I and II. Early work on haemagglutinin (HA), an influenza envelope glycoprotein, extensively characterised the folding pathway of this protein. Experiments using the bicyclic iminosugar castanospermine (CAST) demonstrated that calnexin was required for efficient folding of HA and that the association with calnexin was dependent on the nature of the glycan [2,15,27,28]. The misfolded states of other viral proteins, such as HIV [29], have also been characterised and direct links observed between iminosugars, misfolding and the antiviral effect. However, in the majority of the cases the evidence for this mechanism of action is less direct; addition of iminosugar leads to impaired virion secretion or secretion of non-infectious virions. Here, we describe the calnexin cycle in more detail and briefly touch on other potential mechanisms of action which may be relevant in some viruses.

N-linked glycoproteins are co-translationally modified at the luminal face of the ER through en bloc addition of one or more precursor glycans in the form of Glc3Man9GlcNAc2 by oligosaccharyltransferase [30]. The glycans are recognised and processed by a variety of ER- and Golgi-resident factors that assist protein folding and assembly, mediate flow of secretory cargo and trigger ER-associated degradation (ERAD) [31,32]. Sequential cleavage of the two terminal glucose residues is important for interaction of the nascent polypeptide chain with calnexin, which forms a core part of the ER quality control (ERQC) mechanism [2,15,33,34]. ER α-glu I and α-glu II are the gatekeepers for the calnexin cycle, with binding to ERQC components dependent on the glycoform that the nascent polypeptide retains. ER α-glu I cleaves the terminal glucose residue of the N-linked glycan to give a Glc2Man9GlcNAc2 species. This diglucosylated glycan can be specifically bound by malectin, a membrane-bound ER-resident lectin [35]. Expression of malectin is induced by the unfolded protein response [36], and the protein is proposed to preferentially associate with non-native conformers of folding glycoproteins [37]. The glycan-bound form of malectin potentially associates with the translocon-associated oligosaccharyl transferase acting as an early pathway misfolding sensor [38]. Cleavage of the second glucose residue by α-glu II results in Glc1Man9GlcNAc2, which competes for binding with calnexin/calreticulin and α-glu II [33]. Binding by calnexin retains the protein in the ER where it can interact with chaperones such as binding immunoglobulin protein (BiP) and protein disulfide-isomerase (PDI) [34]. Binding to α-glu II results in cleavage of the third glucose residue after which there are several possible outcomes. If the protein is correctly folded, it can move to the Golgi apparatus for further processing of the glycans. If the protein is misfolded, this may be recognised by UDP-glucose:glycoprotein glucosyl transferase (UGGT), which reglucosylates the glycan such that the protein is once again a substrate for calnexin [39,40]; alternatively, the protein may encounter an α-mannosidase which removes a specific terminal mannose residue targeting the protein for degradation (Figure 2) [41,42].

The presence of large quantities of misfolded proteins will trigger ERAD [32]. This pathway targets misfolded proteins for translocation from the ER into the cytosol, ubiquitination and subsequent hydrolysis by the proteasome. The ERAD targeting presumably occurs through a variety of mechanisms, depending on the nature of the substrate as well as the localisation of the misfolded region within the protein. Glycoproteins degraded through ERAD have their glycan portion released prior to the proteasomal destruction in the cytosol by a peptide:N-glycanase (PNGase) [4345]. As a result, free oligosaccharides (FOS) are produced. These FOS are an excellent biomarker for ERAD and provide both qualitative and quantitative insight into glycoprotein degradation in the ER on a cellular level.

The FOS produced by PNGase initially carry two GlcNAc residues at the reducing terminus. The cytosolic pathway for FOS catabolism in normal conditions involves rapid processing by endo-β-N-acetylglucosaminidase (ENGase), which removes the terminal GlcNAc residue, resulting in a mono-GlcNAc form of FOS [46,47]. This is trimmed by a cytosolic neutral α-mannosidase (NAM, MAN2C1) and subsequently transported into the lysosome for further hydrolysis to monosaccharides [4850]. Inhibition of α-glu I and II leads to the presence of glucosylated FOS [51,52]. These are processed in the cytosol to the final form of Glc3Man5GlcNAc1 (and slowly to Glc3Man4GlcNAc1; Figure 3). However, glucosylated FOS are unable to access the lysosome, resulting in their cytosolic accumulation [50]. The use of the ER glucosidase iminosugar inhibitor N-butyl-deoxynojirimycin (NB-DNJ) has demonstrated that the generation of glucosylated FOS is both dose- and treatment time-dependent. Analysing the level of glucosylated FOS in the cytosol is a valuable test for novel iminosugars in a cellular context [53]. This is particularly pertinent, as although there are in vitro assays for α-glucosidase inhibition, these do not address the question of cellular uptake. Entry of iminosugars into the ER needs to be achieved and demonstrated for these compounds to be developed for clinical trials.

Using FOS to probe iminosugar mechanism of action

The calnexin cycle provides a simplistic picture of the current targets involved in inhibiting glycoprotein folding — namely the enzymes α-glu I and α-glu II. The sequential removal of glucose residues, the flux through the ER and the multiple equilibria (between enzymes, inhibitors and competing substrates) involved in the processing of the glycoprotein make it difficult to predict the extent to which each of the two enzymes will be inhibited for any given concentration of iminosugar. We might have expected, for example, that inhibition of α-glu I, which iminosugars such as NB-DNJ inhibit at least 10 times more strongly than α-glu II (Table 1), would dominate. However, FOS analysis shows that, at physiologically significant iminosugar concentrations, where an antiviral effect is observed, the result of all the competing reactions and processes is that inhibition of the α-glu II enzyme at the mono-glucosylated stage is key, and that higher concentrations of iminosugars are needed to inhibit predominately α-glu I [53].

The FOS assay carried out on cells incubated with NB-DNJ revealed an initial build-up of mono-glucosylated FOS at short treatment periods and low concentrations of iminosugar, before triglucosylated FOS are generated [53]. This reveals the potent cellular inhibition of α-glu II, compared with α-glu I, and contrasts markedly with the inhibition of these enzymes using in vitro assays, where, on the basis of IC50 values, NB-DNJ is 100 times more efficient at inhibiting α-glu I than II. This result is mirrored by more potent α-glu I inhibitors such as NAP-DNJ, which inhibits at low nanomolar concentrations (Table 1) [54]. The very small amount of diglucosylated FOS produced by iminosugar inhibition in the cell demonstrates that iminosugars are relatively poor at preventing the removal of the first α1,2-linked glucose. This reflects the kinetics of α-glu I and II action in the ER, which results in the efficient and rapid hydrolysis of tri- and di-glucosylated glycans to mono-glucosylated glycans, allowing interaction with calnexin/calreticulin chaperones.

The longer half-life of mono-glucosylated glycans, compared with that of tri- and di-glucosylated species, is due to the slower hydrolysis rate of the proximal glucose residue by α-glu II [55]. This contributes to a more favourable environment for iminosugar inhibition of this step. Removal of the first glucose residue by α-glu I and the second glucose residue by α-glu II in cultured cells probably occurs close to their limiting rates (Vmax), where addition of a competitive inhibitor has a limited effect on the observed rate. Removal of the third glucose residue by α-glu II is much slower, suggesting that this rate is not close to its Vmax, and under such conditions a competitive inhibitor has a greater effect on the rate. The calnexin molecule acts as an anchor, while chaperones, such as BiP and PDI, help with the folding process. Since α-glucosidase II and calnexin both compete for the same substrate in the ER, Glc1Man9GlcNac2-protein, if α-glu II is unable to hydrolyse the substrate bound to calnexin, the presence of calnexin will significantly reduce the free substrate concentration, hence reducing the rate. Alternatively, if α-glu II was able to hydrolyse the substrate bound to calnexin (although such a scenario is difficult to envisage), then the presence of calnexin is likely to change the Km. An increase in Km would also result in a reduced rate. Both possibilities have the same functional outcome: the rate of removal of the proximal glucose residue is reduced in cells and hence is more sensitive to the presence of a competitive inhibitor, resulting in a greater accumulation of mono-glucosylated glycans. The retention of the mono-glucosylated form by inhibition of α-glu II by an iminosugar will result in the glycoprotein failing to pass the quality control and it being targeted for degradation in the ER. It is important to note that not all glycoproteins require the calnexin cycle to fold correctly. The Glc1Man9GlcNAc2 glycan exists in a dynamic equilibrium between two structures, a major and a minor conformer for the Glcα1–3Manα linkage [56]. Modelling of the binding of Glc1Man9GlcNAc2 to calnexin suggests that it is the minor conformer that is recognised by calnexin. This may be one of the mechanisms for controlling the rate of recruitment of proteins into the calnexin/calreticulin chaperone system and enabling proteins that do not use this particular quality control mechanism for folding to bypass the system.

Inhibition of glycolipid processing

Although it is clear that inhibition of ER α-glu I and/or II has an antiviral effect against many viruses, this may not be the whole story. Monocyclic compounds such as 1-deoxynojirimycin (DNJ) and its derivatives are also inhibitors of glycolipid processing enzymes [52]. Glycolipids play an important role in the life-cycle of many viruses; neutral sphingolipids are thought to be important for the binding of DENV to mammalian and mosquito cell surfaces [57,58], whilst glycosphingolipids are a major component of the lipid rafts necessary for the replication of HCV [59,60]. It is therefore possible that the antiviral effects of iminosugars against these viruses arise from inhibition of glycolipid processing in addition to inhibiting glycoprotein folding. The extent to which inhibition of glycolipid processing contributes to antiviral effects will almost certainly be virus-dependent. Recently, Sayce et al. demonstrated that, in the case of DENV, the antiviral effects arise via inhibition of glycoprotein folding and that iminosugar effects on glycolipid processing do not contribute to the antiviral effect [61,62]. For HCV and other viruses the picture is less clear and investigations in this area are ongoing.

It is clear that both α-glucosidases are important antiviral targets. Recent structural biology approaches have attempted to examine their structures to both understand the mechanism of inhibition and provide insights into the design and evaluation of more selective and potent inhibitors [6365]. Inhibition is based on a DNJ structure, which has glucose stereochemistry, and hence, with the abundance of glucose usage in the cell, the chances of multiple binding partners for the drugs are high. Indeed, DNJ compounds with alkyl chains have shown efficacy against enzymes including ER α-glucosidases, gut α-glucosidases, ceramide glucosyl transferase Β and glucocerebrosidases (GBA 1, 2 and 3). The 2.04 Å crystal structure of ER α-glu I determined in 2013 provides clear indication of the catalytic residues, but unfortunately the crystal packing prohibits experimental active-site investigations due to occlusion of the active site by a His6 purification tag from a crystal contact [63]. More recently, Caputo et al. [64] have solved the structure of mouse α-glu II (Mmα-GluII) in the presence of a series of substrate analogues, inhibitors and products (Figure 4). When compared with the structure of an intestinal α-glucosidase, the structure has revealed some crucial differences including an ‘exclusion loop’ unique to α-glu II that hangs over the active site. Generating compounds that bind to the exclusion loop region may therefore lead to increased specificity for ER α-glu II over the intestinal α-glucosidases. In addition, crystal structures with inhibitors gave insight into potential methods for designing inhibitors with increased affinity. Alkylated DNJ derivatives, for example, occupy the site of the terminal glucose residue as expected, but the extent to which they occupy the binding sites for the adjacent sugar moieties depends upon the length of the alkyl chain. NB-DNJ, with its four carbon alkyl chain, occupies the +1 sugar-binding site, adjacent to the terminal glucose site, whilst MON-DNJ with its much longer chain extends towards the +2 sugar-binding site. The co-crystal structure with MON-DNJ suggests that the aromatic residues forming a ring between the +1 and +2 site would be good targets for inhibitors with increased potency (Figure 4C,D).

Iminosugars show antiviral activity against viruses from a wide range of different families, thus it appears that glycoproteins from many viruses are dependent upon the calnexin cycle for proper folding. There is, however, indirect evidence that the extent of misfolding caused by iminosugar inhibition of α-glucosidases varies between viruses. For some viruses, such as HIV, the level of secreted virus decreases by only a small amount in the presence of iminosugars, but the secreted viral particles have greatly reduced infectivity [29]. This implies that the degree of misfolding is small enough to escape detection by the cellular unfolded protein response but large enough to impair function. In contrast, for other viruses (including DENV, Hepatitis B virus and bovine viral diarrhea virus) the presence of iminosugars leads to greatly reduced secretion of virions [1,19,66]. This suggests that the misfolding of the glycoproteins in this case is more severe. In the case of Hepatitis B, at high doses of iminosugars (where α-glu I is inhibited and viral glycoproteins with the unmodified Glc3Man9GlcNAc2 glycoform will predominate) a considerable fraction of the viral glycoproteins are retained within the cell. These glycoproteins appear as aggregates and have an intracellular half-life in excess of 24 h. It was hypothesised that one particular viral glycoprotein (M) when in the triglucosylated state may act as a kind of ‘poison pill' in the ER preventing the secretion of the virus. In this case a viral protein acts as a ‘dominant negative' poison of viral secretion and may itself be considered the antiviral drug. The long intracellular half-life of the ‘poison pill’ may be due to aberrant trafficking between the ER and Golgi that has recently been observed for glycoproteins retaining terminal glucose residues on their glycans. This additional mechanism could lead to a reservoir of unfolded protein interfering with the ERQC. These are areas that need to be investigated further to understand the structural requirements for viral assembly, secretion and fusion to host cells in an iminosugar protein folding compromised context.

Currently, we cannot yet predict which glycoproteins will be secreted and which degraded in the presence of an iminosugar, and it remains a challenge to identify whether inhibition of ER α-glu I versus α-glu II will affect the extent of misfolding observed.

Naturally occurring iminosugars including DNJ and CAST have been used as the starting point for medicinal chemistry strategies aimed at antiviral drug discovery. Compounds modified from these starting materials have been, or are currently being, evaluated in clinical trials against HIV, HCV and DENV [6771].

The n-butylated form of DNJ (NB-DNJ) and celgosivir, a prodrug of CAST, were both evaluated in clinical trials against HIV; however, neither proceeded further than phase II [67,68]. In the case of NB-DNJ, although some effects on viraemia were observed, it proved impossible to maintain therapeutic concentrations of the drug in serum. Observations that the concentrations of NB-DNJ necessary to inhibit the ER α-glucosidases in whole-cell assays greatly exceed that necessary to achieve the same inhibition of isolated enzymes suggested that difficulties in penetration of the drug into the ER were responsible for these poor results. These trials did, however, demonstrate that both DNJ and celgosivir are well tolerated by patients. The principal side effect of DNJ-derived iminosugars is osmotic diarrhoea, which arises due to the inhibition of the gut α-glucosidases [68,72]. This side effect is reversible and can be controlled by diet; however, improvements to limit the unwanted effect would be desirable, particularly in the development of iminosugar antivirals to treat chronic viral infections.

The current most promising targets for iminosugar antivirals appear to be DENV, for which clinical trials are in process [70,71], and influenza, where promising results have been observed in preclinical studies [21,73]. DENV is a mosquito-borne disease estimated to infect up to 390 million people per annum [74]. Many of these infections have no clinical manifestation, whilst others require hospitalisation, and a small percentage develop into dengue haemorrhagic fever, which can be fatal [75]. Both celgosivir and the DNJ-derivative UV-4 (MON-DNJ) are currently in clinical trials against DENV. UV-4 is in a phase I clinical trial, with a study evaluating the pharmacokinetics of the hydrochloride salt of UV-4 (UV-4B) administered as a multiple ascending dose [71]. Celgosivir has recently been approved for a phase II clinical trial [70]. In a proof of concept phase 1b study in which 50 patients were recruited, administration of celgosivir did not appear to reduce the viral load or fever burden in patients with uncomplicated dengue fever [76]. However, further investigations using a mouse model suggest that the dosing regime may be critical for efficacy. Mouse studies carried out prior to the phase 1b study suggested that a twice-daily treatment regime would be sufficient to reduce viraemia [77]; however, work carried out in light of the results from the phase 1b clinical trial suggests that this treatment regime is inadequate when the treatment is begun at the peak of viraemia [78]. More recent work suggests that a four-time daily regime is likely to be more effective against established infections and this will be investigated in the upcoming trial. A range of iminosugars, including NB-DNJ, NN-DNJ and UV-4B, have been tested for efficacy against the influenza virus [20,21,73]. Intriguingly, the antiviral efficacy observed varies between different influenza strains and subtypes in addition to varying with the specific iminosugar inhibitor tested. In a recent study Warfield et al. demonstrated that, at high concentrations, UV4-B shows activity against both influenza A and influenza B strains in cell culture, and that it gave a significant survival benefit against influenza A H1N1 and H3N2, and influenza B/Sichuan/379/99 in lethal mouse models. Future studies on the efficacy against influenza in humans are planned [21].

Our understanding of the mechanism of action of iminosugar antivirals has developed enormously since DNJ was first reported to have an antiviral effect against HIV. Medicinal chemistry approaches have improved the efficacy and pharmacokinetics of natural iminosugars, and current compounds represent good prospects for iminosugar antivirals reaching the clinic. However, medicinal chemists are already leading the way towards the ‘next generation’ of iminosugars. The compounds currently under investigation are well tolerated by patients and offer hope against current and future viral challenges; however, it is always desirable to limit off-target effects and increase ER uptake. The structures of mammalian ER α-glucosidases, together with those of off-target α-glucosidases, have the potential to facilitate rational design of new inhibitors with increased specificity for the ER-resident enzyme(s).BST-2016-0182CTB1 

BiP

binding immunoglobulin protein

CAST

castanospermine

DENV

dengue virus

ENGase

endo-β-N-acetylglucosaminidase

ER

endoplasmic reticulum

ERAD

ER-associated degradation

ERQC

ER quality control

FOS

free oligosaccharides

HA

haemagglutinin

HCV

hepatitis C virus

MON-DNJ

N-(9′-methoxynonyl)-1-deoxynojirimycin

NAP-DNJ

N-(6′-4″-azido-2″-nitrophenylamino) hexyl-1-deoxynojirimycin

NB-DNJ

N-butyl-deoxynojirimycin

NN-DNJ

N-nonyl-deoxynojirimycin

PDI

protein disulfide-isomerase

PNGase

peptide:N-glycanase

UGGT

UDP-glucose:glycoprotein glucosyl transferase

The Authors declare that there are no competing interests associated with the manuscript.

1
Block
,
T.M.
,
Lu
,
X.
,
Platt
,
F.M.
,
Foster
,
G.R.
,
Gerlich
,
W.H.
,
Blumberg
,
B.S.
et al
(
1994
)
Secretion of human hepatitis B virus is inhibited by the imino sugar N-butyldeoxynojirimycin
.
Proc. Natl Acad. Sci. U.S.A.
91
,
2235
2239
doi:
2
Hammond
,
C.
,
Braakman
,
I.
and
Helenius
,
A.
(
1994
)
Role of N-linked oligosaccharide recognition, glucose trimming, and calnexin in glycoprotein folding and quality control
.
Proc. Natl Acad. Sci. U.S.A.
91
,
913
917
doi:
3
Mehta
,
A.
,
Lu
,
X.
,
Block
,
T.M.
,
Blumberg
,
B.S.
and
Dwek
,
R.A.
(
1997
)
Hepatitis B virus (HBV) envelope glycoproteins vary drastically in their sensitivity to glycan processing: evidence that alteration of a single N-linked glycosylation site can regulate HBV secretion
.
Proc. Natl Acad. Sci. U.S.A.
94
,
1822
1827
doi:
4
Dwek
,
R.A.
,
Butters
,
T.D.
,
Platt
,
F.M.
and
Zitzmann
,
N.
(
2002
)
Targeting glycosylation as a therapeutic approach
.
Nat. Rev. Drug Discovery
1
,
65
75
doi:
5
Chang
,
J.
,
Guo
,
J.-T.
,
Du
,
Y.
and
Block
,
T.
(
2013
)
Imino sugar glucosidase inhibitors as broadly active anti-filovirus agents
.
Emerg. Microbes Infect.
2
,
e77
doi:
6
Scott
,
L.J.
(
2016
)
Tetravalent dengue vaccine: a review in the prevention of dengue disease
.
Drugs
76
,
1301
1312
doi:
7
Ahmed
,
A.
and
Felmlee
,
D.J.
(
2015
)
Mechanisms of hepatitis C viral resistance to direct acting antivirals
.
Viruses
7
,
6716
6729
doi:
8
Iyidogan
,
P.
and
Anderson
,
K.S.
(
2014
)
Current perspectives on HIV-1 antiretroviral drug resistance
.
Viruses
6
,
4095
4139
doi:
9
Paulsen
,
H.
(
1966
)
Carbohydrates containing nitrogen or sulfur in the ‘hemiacetal’ ring
.
Angew. Chem. Int. Ed. Engl.
5
,
495
510
doi:
10
Platt
,
F.M.
,
Neises
,
G.R.
,
Dwek
,
R.A.
and
Butters
,
T.D.
(
1994
)
N-butyldeoxynojirimycin is a novel inhibitor of glycolipid biosynthesis
.
J. Biol. Chem.
269
,
8362
8365
PMID:
[PubMed]
11
Cox
,
T.
,
Lachmann
,
R.
,
Hollak
,
C.
,
Aerts
,
J.
,
van Weely
,
S.
,
Hrebícek
,
M.
et al
(
2000
)
Novel oral treatment of Gaucher's disease with N-butyldeoxynojirimycin (OGT 918) to decrease substrate biosynthesis
.
Lancet
355
,
1481
1485
doi:
12
Joubert
,
P.H.
,
Venter
,
C.P.
,
Joubert
,
H.F.
and
Hillebrand
,
I.
(
1985
)
The effect of a 1-deoxynojirimycin derivative on post-prandial blood glucose and insulin levels in healthy black and white volunteers
.
Eur. J. Clin. Pharmacol.
28
,
705
708
doi:
13
Shailubhai
,
K.
,
Pukazhenthi
,
B.S.
,
Saxena
,
E.S.
,
Varma
,
G.M.
and
Vijay
,
I.K.
(
1991
)
Glucosidase I, a transmembrane endoplasmic reticular glycoprotein with a luminal catalytic domain
.
J. Biol. Chem.
266
,
16587
16593
PMID:
[PubMed]
14
Karlsson
,
G.B.
,
Butters
,
T.D.
,
Dwek
,
R.A.
and
Platt
,
F.M.
(
1993
)
Effects of the imino sugar N-butyldeoxynojirimycin on the N-glycosylation of recombinant gp120
.
J. Biol. Chem.
268
,
570
576
PMID:
[PubMed]
15
Chen
,
W.
,
Helenius
,
J.
,
Braakman
,
I.
and
Helenius
,
A.
(
1995
)
Cotranslational folding and calnexin binding during glycoprotein synthesis
.
Proc. Natl Acad. Sci. U.S.A.
92
,
6229
6233
doi:
16
Caramelo
,
J.J.
and
Parodi
,
A.J.
(
2008
)
Getting in and out from calnexin/calreticulin cycles
.
J. Biol. Chem.
283
,
10221
10225
doi:
17
Eisenberg
,
R.J.
,
Atanasiu
,
D.
,
Cairns
,
T.M.
,
Gallagher
,
J.R.
,
Krummenacher
,
C.
and
Cohen
,
G.H.
(
2012
)
Herpes virus fusion and entry: a story with many characters
.
Viruses
4
,
800
832
doi:
18
Wu
,
S.-F.
,
Lee
,
C.-J.
,
Liao
,
C.-L.
,
Dwek
,
R.A.
,
Zitzmann
,
N.
and
Lin
,
Y.-L.
(
2002
)
Antiviral effects of an iminosugar derivative on flavivirus infections
.
J. Virol.
76
,
3596
3604
doi:
19
Courageot
,
M.-P.
,
Frenkiel
,
M.-P.
,
Dos Santos
,
C.D.
,
Deubel
,
V.
and
Desprès
,
P.
(
2000
)
Alpha-Glucosidase inhibitors reduce dengue virus production by affecting the initial steps of virion morphogenesis in the endoplasmic reticulum
.
J. Virol.
74
,
564
572
doi:
20
Hussain
,
S.
,
Miller
,
J.L.
,
Harvey
,
D.J.
,
Gu
,
Y.
,
Rosenthal
,
P.B.
,
Zitzmann
,
N.
et al
(
2015
)
Strain-specific antiviral activity of iminosugars against human influenza A viruses
.
J. Antimicrob. Chemother.
70
,
136
152
doi:
21
Warfield
,
K.L.
,
Barnard
,
D.L.
,
Enterlein
,
S.G.
,
Smee
,
D.F.
,
Khaliq
,
M.
,
Sampath
,
A.
et al
(
2016
)
The iminosugar UV-4 is a broad inhibitor of influenza A and B viruses ex vivo and in mice
.
Viruses
8
,
71
doi:
22
Chapel
,
C.
,
Garcia
,
C.
,
Bartosch
,
B.
,
Roingeard
,
P.
,
Zitzmann
,
N.
,
Cosset
,
F.-L.
et al
(
2007
)
Reduction of the infectivity of hepatitis C virus pseudoparticles by incorporation of misfolded glycoproteins induced by glucosidase inhibitors
.
J. Gen. Virol.
88
,
1133
1143
doi:
23
Steinmann
,
E.
,
Whitfield
,
T.
,
Kallis
,
S.
,
Dwek
,
R.A.
,
Zitzmann
,
N.
,
Pietschmann
,
T.
et al
(
2007
)
Antiviral effects of amantadine and iminosugar derivatives against hepatitis C virus
.
Hepatology
46
,
330
338
doi:
24
Gruters
,
R.A.
,
Neefjes
,
J.J.
,
Tersmette
,
M.
,
de Goede
,
R.E.Y.
,
Tulp
,
A.
,
Huisman
,
H.G.
et al
(
1987
)
Interference with HIV-induced syncytium formation and viral infectivity by inhibitors of trimming glucosidase
.
Nature
330
,
74
77
doi:
25
Fleet
,
G.W.J.
,
Karpas
,
A.
,
Dwek
,
R.A.
,
Fellows
,
L.E.
,
Tyms
,
A.S.
,
Petursson
,
S.
et al
(
1988
)
Inhibition of HIV replication by amino-sugar derivatives
.
FEBS Lett.
237
,
128
132
doi:
26
Sadat
,
M.A.
,
Moir
,
S.
,
Chun
,
T.-W.
,
Lusso
,
P.
,
Kaplan
,
G.
,
Wolfe
,
L.
et al
(
2014
)
Glycosylation, hypogammaglobulinemia, and resistance to viral infections
.
N. Engl. J. Med.
370
,
1615
1625
doi:
27
Tatu
,
U.
,
Hammond
,
C.
and
Helenius
,
A.
(
1995
)
Folding and oligomerization of influenza hemagglutinin in the ER and the intermediate compartment
.
EMBO J.
14
,
1340
1348
PMID:
[PubMed]
28
Hebert
,
D.N.
,
Foelimer
,
B.
and
Helenius
,
A.
(
1996
)
Calnexin and calreticulin promote folding, delay oligomerization and suppress degradation of influenza hemagglutinin in microsomes
.
EMBO J.
15
,
2961
2968
PMID:
[PubMed]
29
Fischer
,
P.B.
,
Karlsson
,
G.B.
,
Butters
,
T.D.
,
Dwek
,
R.A.
and
Platt
,
F.M.
(
1996
)
N-butyldeoxynojirimycin-mediated inhibition of human immunodeficiency virus entry correlates with changes in antibody recognition of the V1/V2 region of gp120
.
J. Virol.
70
,
7143
7152
PMID:
[PubMed]
30
Kelleher
,
D.J.
and
Gilmore
,
R.
(
2006
)
An evolving view of the eukaryotic oligosaccharyltransferase
.
Glycobiology
16
,
47R
62R
doi:
31
Xu
,
C.
and
Ng
,
D.T.W.
(
2015
)
Glycosylation-directed quality control of protein folding
.
Nat. Rev. Mol. Cell Biol.
16
,
742
752
doi:
32
Ruggiano
,
A.
,
Foresti
,
O.
and
Carvalho
,
P.
(
2014
)
ER-associated degradation: protein quality control and beyond
.
J. Cell Biol.
204
,
869
879
doi:
33
Hebert
,
D.N.
,
Foellmer
,
B.
and
Helenius
,
A.
(
1995
)
Glucose trimming and reglucosylation determine glycoprotein association with calnexin in the endoplasmic reticulum
.
Cell
81
,
425
433
doi:
34
Hammond
,
C.
and
Helenius
,
A.
(
1994
)
Quality control in the secretory pathway: retention of a misfolded viral membrane glycoprotein involves cycling between the ER, intermediate compartment, and Golgi apparatus
.
J. Cell Biol.
126
,
41
52
doi:
35
Schallus
,
T.
,
Jaeckh
,
C.
,
Fehér
,
K.
,
Palma
,
A.S.
,
Liu
,
Y.
,
Simpson
,
J.C.
et al
(
2008
)
Malectin: a novel carbohydrate-binding protein of the endoplasmic reticulum and a candidate player in the early steps of protein N-glycosylation
.
Mol. Biol. Cell
19
,
3404
3414
doi:
36
Galli
,
C.
,
Bernasconi
,
R.
,
Soldà
,
T.
,
Calanca
,
V.
and
Molinari
,
M.
(
2011
)
Malectin participates in a backup glycoprotein quality control pathway in the mammalian ER
.
PLoS ONE
6
,
e16304
doi:
37
Chen
,
Y.
,
Hu
,
D.
,
Yabe
,
R.
,
Tateno
,
H.
,
Qin
,
S.-Y.
,
Matsumoto
,
N.
et al
(
2011
)
Role of malectin in Glc2Man9GlcNAc2-dependent quality control of α1-antitrypsin
.
Mol. Biol. Cell
22
,
3559
3570
doi:
38
Qin
,
S.-Y.
,
Hu
,
D.
,
Matsumoto
,
K.
,
Takeda
,
K.
,
Matsumoto
,
N.
,
Yamaguchi
,
Y.
et al
(
2012
)
Malectin forms a complex with ribophorin I for enhanced association with misfolded glycoproteins
.
J. Biol. Chem.
287
,
38080
38089
doi:
39
Trombetta
,
S.E.
and
Parodi
,
A.J.
(
1992
)
Purification to apparent homogeneity and partial characterization of rat liver UDP-glucose:glycoprotein glucosyltransferase
.
J. Biol. Chem.
267
,
9236
9240
PMID:
[PubMed]
40
Taylor
,
S.C.
,
Ferguson
,
A.D.
,
Bergeron
,
J.J.M.
and
Thomas
,
D.Y.
(
2004
)
The ER protein folding sensor UDP-glucose glycoprotein–glucosyltransferase modifies substrates distant to local changes in glycoprotein conformation
.
Nat. Struct. Mol. Biol.
11
,
128
134
doi:
41
Weng
,
S.
and
Spiro
,
R.G.
(
1993
)
Demonstration that a kifunensine-resistant α-mannosidase with a unique processing action on N-linked oligosaccharides occurs in rat liver endoplasmic reticulum and various cultured cells
.
J. Biol. Chem.
268
,
25656
25663
PMID:
[PubMed]
42
Tokunaga
,
F.
,
Brostrom
,
C.
,
Koide
,
T.
and
Arvan
,
P.
(
2000
)
Endoplasmic reticulum (ER)-associated degradation of misfolded N-linked glycoproteins is suppressed upon inhibition of ER mannosidase I
.
J. Biol. Chem.
275
,
40757
40764
doi:
43
Weng
,
S.
and
Spiro
,
R.G.
(
1997
)
Demonstration of a peptide:N-glycosidase in the endoplasmic reticulum of rat liver
.
Biochem. J.
322
,
655
661
doi:
44
Suzuki
,
T.
,
Park
,
H.
and
Lennarz
,
W.J.
(
2016
)
Cytoplasmic peptide:N-glycanase (PNGase) in eukaryotic cells: occurrence, primary structure, and potential functions
.
FASEB J.
16
,
635
641
doi:
45
Suzuki
,
T.
,
Kitajima
,
K.
,
Emori
,
Y.
,
Inoue
,
Y.
and
Inoue
,
S.
(
1997
)
Site-specific de-N-glycosylation of diglycosylated ovalbumin in hen oviduct by endogenous peptide:N-glycanase as a quality control system for newly synthesized proteins
.
Proc. Natl Acad. Sci. U.S.A.
94
,
6244
6249
doi:
46
Pierce
,
R.J.
,
Spik
,
G.
and
Montreuil
,
J.
(
1979
)
Cytosolic location of an endo-N-acetyl-β-d-glucosaminidase activity in rat liver and kidney
.
Biochem. J.
180
,
673
676
doi:
47
Kato
,
T.
,
Fujita
,
K.
,
Takeuchi
,
M.
,
Kobayashi
,
K.
,
Natsuka
,
S.
,
Ikura
,
K.
et al
(
2002
)
Identification of an endo-β-N-acetylglucosaminidase gene in Caenorhabditis elegans and its expression in Escherichia coli.
Glycobiology
12
,
581
587
PMID:
[PubMed]
48
Shoup
,
V.A.
and
Touster
,
O.
(
1976
)
Purification and characterization of the α-d-mannosidase of rat liver cytosol
.
J. Biol. Chem.
251
,
3845
3852
PMID:
[PubMed]
49
Suzuki
,
T.
,
Hara
,
I.
,
Nakano
,
M.
,
Shigeta
,
M.
,
Nakagawa
,
T.
,
Kondo
,
A.
et al
(
2006
)
Man2c1, an α-mannosidase, is involved in the trimming of free oligosaccharides in the cytosol
.
Biochem. J.
400
,
33
41
doi:
50
Saint-Pol
,
A.
,
Codogno
,
P.
and
Moore
,
S.E.H.
(
1999
)
Cytosol-to-lysosome transport of free polymannose-type oligosaccharides: kinetic and specificity studies using rat liver lysosomes
.
J. Biol. Chem.
274
,
13547
13555
doi:
51
Durrant
,
C.
and
Moore
,
S.E.H.
(
2002
)
Perturbation of free oligosaccharide trafficking in endoplasmic reticulum glucosidase I-deficient and castanospermine-treated cells
.
Biochem. J.
365
,
239
247
doi:
52
Mellor
,
H.R.
,
Neville
,
D.C.A.
,
Harvey
,
D.J.
,
Platt
,
F.M.
,
Dwek
,
R.A.
and
Butters
,
T.D.
(
2004
)
Cellular effects of deoxynojirimycin analogues: inhibition of N-linked oligosaccharide processing and generation of free glucosylated oligosaccharides
.
Biochem. J.
381
,
867
875
doi:
53
Alonzi
,
D.S.
,
Neville
,
D.C.A.
,
Lachmann
,
R.H.
,
Dwek
,
R.A.
and
Butters
,
T.D.
(
2008
)
Glucosylated free oligosaccharides are biomarkers of endoplasmic-reticulum α-glucosidase inhibition
.
Biochem. J.
409
,
571
580
doi:
54
Rawlings
,
A.J.
,
Lomas
,
H.
,
Pilling
,
A.W.
,
Lee
,
M.J.-R.
,
Alonzi
,
D.S.
,
Rountree
,
J.S.S.
et al
(
2009
)
Synthesis and biological characterisation of novel N-alkyl-deoxynojirimycin α-glucosidase inhibitors
.
ChemBioChem
10
,
1101
1105
doi:
55
Grinna
,
L.S.
and
Robbins
,
P.W.
(
1980
)
Substrate specificities of rat liver microsomal glucosidases which process glycoproteins
.
J. Biol. Chem.
255
,
2255
2258
PMID:
[PubMed]
56
Mackeen
,
M.M.
,
Almond
,
A.
,
Deschamps
,
M.
,
Cumpstey
,
I.
,
Fairbanks
,
A.J.
,
Tsang
,
C.
et al
(
2009
)
The conformational properties of the Glc3Man unit suggest conformational biasing within the chaperone-assisted glycoprotein folding pathway
.
J. Mol. Biol.
387
,
335
347
doi:
57
Wichit
,
S.
,
Jittmittraphap
,
A.
,
Hidari
,
K.I.P.J.
,
Thaisomboonsuk
,
B.
,
Petmitr
,
S.
,
Ubol
,
S.
et al
(
2011
)
Dengue virus type 2 recognizes the carbohydrate moiety of neutral glycosphingolipids in mammalian and mosquito cells
.
Microbiol. Immunol.
55
,
135
140
doi:
58
Perera
,
R.
,
Riley
,
C.
,
Isaac
,
G.
,
Hopf-Jannasch
,
A.S.
,
Moore
,
R.J.
,
Weitz
,
K.W.
et al
(
2012
)
Dengue virus infection perturbs lipid homeostasis in infected mosquito cells
.
PLoS Pathog.
8
,
e1002584
doi:
59
Kapadia
,
S.B.
,
Barth
,
H.
,
Baumert
,
T.
,
McKeating
,
J.A.
and
Chisari
,
F.V.
(
2007
)
Initiation of hepatitis C virus infection is dependent on cholesterol and cooperativity between CD81 and scavenger receptor B type I
.
J. Virol.
81
,
374
383
doi:
60
Voisset
,
C.
,
Lavie
,
M.
,
Helle
,
F.
,
De Beeck
,
A.O.
,
Bilheu
,
A.
,
Bertrand-Michel
,
J.
et al
(
2008
)
Ceramide enrichment of the plasma membrane induces CD81 internalization and inhibits hepatitis C virus entry
.
Cell. Microbiol.
10
,
606
617
doi:
61
Sayce
,
A.C.
,
Alonzi
,
D.S.
,
Killingbeck
,
S.S.
,
Tyrrell
,
B.E.
,
Hill
,
M.L.
,
Caputo
,
A.T.
et al
(
2016
)
Iminosugars inhibit dengue virus production via inhibition of ER alpha-glucosidases — not glycolipid processing enzymes
.
PLoS Negl. Trop. Dis.
10
,
e0004524
doi:
62
Warfield
,
K.L.
,
Plummer
,
E.M.
,
Sayce
,
A.C.
,
Alonzi
,
D.S.
,
Tang
,
W.
,
Tyrrell
,
B.E.
et al
(
2016
)
Inhibition of endoplasmic reticulum glucosidases is required for in vitro and in vivo dengue antiviral activity by the iminosugar UV-4
.
Antiviral Res.
129
,
93
98
doi:
63
Barker
,
M.K.
and
Rose
,
D.R.
(
2013
)
Specificity of processing α-glucosidase I is guided by the substrate conformation: crystallographic and in silico studies
.
J. Biol. Chem.
288
,
13563
13574
doi:
64
Caputo
,
A.T.
,
Alonzi
,
D.S.
,
Marti
,
L.
,
Reca
,
I.-B.
,
Kiappes
,
J.L.
,
Struwe
,
W.B.
et al
(
2016
)
Structures of mammalian ER α-glucosidase II capture the binding modes of broad-spectrum iminosugar antivirals
.
Proc. Natl Acad. Sci. U.S.A.
113
,
E4630
–E4638
doi:
65
Satoh
,
T.
,
Toshimori
,
T.
,
Yan
,
G.
,
Yamaguchi
,
T.
and
Kato
,
K.
(
2016
)
Structural basis for two-step glucose trimming by glucosidase II involved in ER glycoprotein quality control
.
Sci. Rep.
6
,
20575
doi:
66
Zitzmann
,
N.
,
Mehta
,
A.S.
,
Carrouee
,
S.
,
Butters
,
T.D.
,
Platt
,
F.M.
,
McCauley
,
J.
et al
(
1999
)
Imino sugars inhibit the formation and secretion of bovine viral diarrhea virus, a pestivirus model of hepatitis C virus: implications for the development of broad spectrum anti-hepatitis virus agents
.
Proc. Natl Acad. Sci. U.S.A.
96
,
11878
11882
doi:
67
Hoechst Marion Roussel
. (
1996
)
A randomized, double-blind active-controlled, dose-ranging study of the safety and efficacy of chronically administered MDL 28,574A in the treatment of HIV-infected patients
.
NLM identifier: NCT00002151
68
Fischl
,
M.A.
,
Resnick
,
L.
,
Coombs
,
R.
,
Kremer
,
A.B.
,
Pottage
, Jr,
J.C.
,
Fass
,
R.J.
et al
(
1994
)
The safety and efficacy of combination N-butyl-deoxynojirimycin (SC-48334) and zidovudine in patients with HIV-1 infection and 200-500 CD4 cells/mm3
.
J. Acquir. Immune Defic. Syndr.
7
,
139
147
PMID:
[PubMed]
69
BioWest Therapeutics Inc
. (
2006
)
An extension study to evaluate the safety and efficacy of Celgosivir and Peginterferon Alfa-2b, with or without Ribavirin, in patients with chronic Hepatitis C genotype 1 infection
.
NLM identifier: NCT00292084
70
Singapore General Hospital
. (
2016
)
Celgosivir or modipafant as treatment for adult participants with uncomplicated dengue fever in Singapore
.
NLM identifier: NCT02569827
71
Emergent BioSolutions
. (
2016
)
Safety and pharmacokinetics of UV-4B solution administered orally as multiple ascending doses to healthy subjects
.
NLM identifier: NCT02696291
72
Tierney
,
M.
,
Pottage
,
J.
,
Kessler
,
H.
,
Fischl
,
M.A.
,
Richman
,
D.
,
Merigan
,
T.
et al
(
1995
)
The tolerability and pharmacokinetics of N-butyl-deoxynojirimycin in patients with advanced HIV disease (ACTG 100)
.
J. Acquired Immune Defic. Syndr. Hum. Retrovirol.
10
,
549
553
doi:
73
Stavale
,
E.J.
,
Vu
,
H.
,
Sampath
,
A.
,
Ramstedt
,
U.
and
Warfield
,
K.L.
(
2015
)
In vivo therapeutic protection against influenza a (H1N1) oseltamivir-sensitive and resistant viruses by the iminosugar UV-4
.
PLoS ONE
10
,
e0121662
doi:
74
Bhatt
,
S.
,
Gething
,
P.W.
,
Brady
,
O.J.
,
Messina
,
J.P.
,
Farlow
,
A.W.
,
Moyes
,
C.L.
et al
(
2013
)
The global distribution and burden of dengue
.
Nature
496
,
504
507
doi:
75
Halstead
,
S.B.
(
2007
)
Dengue
.
Lancet
370
,
1644
1652
doi:
76
Low
,
J.G.
,
Sung
,
C.
,
Wijaya
,
L.
,
Wei
,
Y.
,
Rathore
,
A.P.S.
,
Watanabe
,
S.
et al
(
2014
)
Efficacy and safety of celgosivir in patients with dengue fever (CELADEN): a phase 1b, randomised, double-blind, placebo-controlled, proof-of-concept trial
.
Lancet Infect. Dis.
14
,
706
715
doi:
77
Watanabe
,
S.
,
Rathore
,
A.P.S.
,
Sung
,
C.
,
Lu
,
F.
,
Khoo
,
Y.M.
,
Connolly
,
J.
et al
(
2012
)
Dose- and schedule-dependent protective efficacy of celgosivir in a lethal mouse model for dengue virus infection informs dosing regimen for a proof of concept clinical trial
.
Antiviral Res.
96
,
32
35
doi:
78
Watanabe
,
S.
,
Chan
,
K.W.-K.
,
Dow
,
G.
,
Ooi
,
E.E.
,
Low
,
J.G.
and
Vasudevan
,
S.G.
(
2016
)
Optimizing celgosivir therapy in mouse models of dengue virus infection of serotypes 1 and 2: the search for a window for potential therapeutic efficacy
.
Antiviral Res.
127
,
10
19
doi:
79
Sousa
,
M.
and
Parodi
,
A.J.
(
1995
)
The molecular basis for the recognition of misfolded glycoproteins by the UDP-Glc:glycoprotein glucosyltransferase
.
EMBO J.
14
,
4196
4203
PMID:
[PubMed]
80
Oda
,
Y.
,
Hosokawa
,
N.
,
Wada
,
I.
and
Nagata
,
K.
(
2003
)
EDEM as an acceptor of terminally misfolded glycoproteins released from calnexin
.
Science
299
,
1394
1397
doi:
81
Hosokawa
,
N.
,
Wada
,
I.
,
Natsuka
,
Y.
and
Nagata
,
K.
(
2006
)
EDEM accelerates ERAD by preventing aberrant dimer formation of misfolded α1-antitrypsin
.
Genes Cells
11
,
465
476
doi:
82
Hirao
,
K.
,
Natsuka
,
Y.
,
Tamura
,
T.
,
Wada
,
I.
,
Morito
,
D.
,
Natuska
,
S.
et al
(
2006
)
EDEM3, a soluble EDEM homolog, enhances glycoprotein endoplasmic reticulum-associated degradation and mannose trimming
.
J. Biol. Chem.
281
,
9650
9658
doi:
83
Olivari
,
S.
,
Cali
,
T.
,
Salo
,
K.E.H.
,
Paganetti
,
P.
,
Ruddock
,
L.W.
and
Molinari
,
M.
(
2006
)
EDEM1 regulates ER-associated degradation by accelerating de-mannosylation of folding-defective polypeptides and by inhibiting their covalent aggregation
.
Biochem. Biophys. Res. Commun.
349
,
1278
1284
doi:
84
Kim
,
W.
,
Spear
,
E.D.
and
Ng
,
D.T.W.
(
2005
)
Yos9p detects and targets misfolded glycoproteins for ER-associated degradation
.
Mol. Cell
19
,
753
764
doi:
85
Buschhorn
,
B.A.
,
Kostova
,
Z.
,
Medicherla
,
B.
and
Wolf
,
D.H.
(
2004
)
A genome-wide screen identifies Yos9p as essential for ER-associated degradation of glycoproteins
.
FEBS Lett.
577
,
422
426
doi:
86
Bhamidipati
,
A.
,
Denic
,
V.
,
Quan
,
E.M.
and
Weissman
,
J.S.
(
2005
)
Exploration of the topological requirements of ERAD identifies Yos9p as a lectin sensor of misfolded glycoproteins in the ER lumen
.
Mol. Cell
19
,
741
751
doi:
87
Christianson
,
J.C.
,
Shaler
,
T.A.
,
Tyler
,
R.E.
and
Kopito
,
R.R.
(
2008
)
OS-9 and GRP94 deliver mutant α1-antitrypsin to the Hrd1–SEL1L ubiquitin ligase complex for ERAD
.
Nat. Cell Biol.
10
,
272
282
doi:
88
Suzuki
,
T.
,
Huang
,
C.
and
Fujihira
,
H.
(
2016
)
The cytoplasmic peptide:N-glycanase (NGLY1) — structure, expression and cellular functions
.
Gene
577
,
1
7
doi:
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).