Abstract
Subsets of protein toxins utilize gangliosides as host receptors. Gangliosides are preferred receptors due to their extracellular localization on the eukaryotic cell and due to their essential nature in host physiology. Glycosphingolipids, including gangliosides, are mediators of signal transduction within and between eukaryotic cells. Protein toxins possess AB structure-function organization, where the A domain encodes a catalytic function for the posttranslational modification of a host macromolecule, including proteins and nucleic acids, and a B domain, which encodes host receptor recognition, including proteins and glycosphingolipids, alone or in combination. Protein toxins use similar strategies to bind glycans by pockets and loops, generally employing hydrogen bonding and aromatic stacking to stabilize interactions with sugars. In some cases, glycan binding facilitates uptake, while in other cases, cross-linking or a second receptor is necessary to stimulate entry. The affinity that protein toxins have for host glycans is necessary for tissue targeting, but not always sufficient to cause disease. In addition to affinity for binding the glycan, the lipid moiety also plays an important role in productive uptake and tissue tropism. Upon endocytosis, the protein toxin must escape to another intracellular compartment or into cytosol to modify a host substrate, modulating host signaling, often resulting in cytotoxic or apoptotic events in the cell, and a unique morbidity for the organism. The study of protein toxins that utilize gangliosides as host receptors has illuminated numerous eukaryotic cellular processes, identified the basis for developing interventions to prevent disease through vaccines and control bacterial diseases through therapies. In addition, subsets of these protein toxins have been utilized as therapeutic agents to treat numerous human inflictions.
Protein toxin AB structure-function organization
AB protein toxins are classified by domain organization (Figure 1). A is the catalytic domain, which posttranslationally modifies a host protein(s), changing host physiology and often causing death or intoxication1. B is the binding domain, which binds a host receptor(s) and may also contain a translocation domain to facilitate delivery of A into the host cytosol1. Inactivation of either the A or B domains results in a nontoxic protein. For single-chain AB toxins, A and B domains are synthesized within a single protein, while for AB5 toxins, A and B domains are synthesized independently and assemble into an oligomer composed of one molecule of A bound to a B pentamer. The A domains of AB5 toxins comprise an A1 catalytic domain and an A2 linker domain. The alpha-helical A2 linker domain inserts into the B pentamer through non-covalent interactions to link the A1 subdomain with the B pentamer to produce the AB5 toxin2.
Figure 1. AB toxin schematic.
AB protein toxins contain two functional properties: a catalytically active A domain and a receptor binding B domain. AB protein toxins include alpha-toxin (Atx), ricin toxin (Rtx), and clostridial tetanus and botulinum neurotoxins (TeNT/BoNT). Alpha-toxin does not contain an interchain disulfide, but contains a ganglioside binding loop between the A and B domains. TeNT/BoNT contain a B domain, which includes a receptor binding domain (HCR) and a translocation domain (HCT). HCT facilitates delivery of A domain, also known as the light chain (LC). AB5 protein toxins include four families: cholera toxin (Ctx), Shiga toxin (Stx), and subtilase cytotoxin (SubAB) contain an A domain that can be divided into A1- and A2- subdomains, and five identical B binding domains. Pertussis toxin (Ptx) contains non-identical B binding components consisting of four homologous subunits S2:S3:S4:S5 in an 1:1:2:1 molecular organization.
Single-chain AB toxins often contain an interchain disulfide bond, while AB5 toxins often contain an intrachain disulfide bond between the A1- and A2- subdomains. In each case, reduction is usually required during the delivery of the A domain into the host cell cytosol and expression of cellular toxicity3–7. AB toxins are typically synthesized in an inactive form and require proteolytic processing to be converted into an active form. Single-chain AB toxins, such as the clostridial botulinum neurotoxins (BoNTs) and tetanus neurotoxin (TeNT), cleave a trypsin-sensitive site located between A and B retaining the interchain disulfide8. AB toxins, such as diphtheria toxin and anthrax toxin, are also proteolyzed by a host protease, furin, for interdomain cleavage or B subunit assembly, respectively9,10. Most AB5 toxins also contain a protease sensitive site, which is proteolyzed by a bacterial- or host- protease10,11. The four AB5 family members; cholera toxin (Ctx), Shiga toxin (Stx), pertussis toxin (Ptx), and subtilase cytotoxin (SubAB), are proteolyzed within the A domain, between the A1 catalytic and A2 linker-subdomains, which are linked by a disulfide bond12–14.
Protein toxin host receptors
AB toxins exploit host receptors to coordinate host cell binding and delivery of A into the host cell cytosol. Members of the BoNT family have dual host receptors, often a glycolipid (ganglioside) and a protein receptor, while TeNT utilizes two gangliosides as receptors15–20. Ctx binds a single class of glycolipid (GM1a) while other members of the Ctx family, the heat labile E. coli enterotoxins (LTx’s) LT-I and LT-IIa-c, bind multiple gangliosides species with varied affinities21,22. Other toxins, such as ricin and Ptx are promiscuous, and bind specific carbohydrate moieties on glycolipids and glycoproteins23,24.
The specificity and affinity of protein toxins for host receptors is often assessed by in vitro biochemical and cell biological techniques, including solid-phase assays, surface plasmon resonance, receptor loading, and receptor depletion. Additional in vivo gene knockout studies confirm receptor specificity and tissue tropism. Ultimately, protein toxin action requires a receptor, a productive pathway to traffic the A domain to the target substrate, and the presence of the target substrate. This review focuses on protein toxins that utilize a subset of glycosphingolipids, gangliosides, and related carbohydrate moieties as host cell receptors.
Protein toxin entry into cells
Protein toxins enter cells by multiple pathways. Most AB toxins bind a cell surface host receptor and enter via receptor-mediated endocytosis. Early endosomal entry involves clathrin-dependent and clathrin-independent pathways such as bulk uptake, caveolae, and micropinocytosis. AB5 toxins Ctx, LTx, Stx, and SubAB are associated with both clathrin-dependent and clathrin-independent pathways depending on the host cell type25–28. Following uptake, the catalytic domain escapes the early endosome upon endosome acidification or retrograde traffics to the endoplasmic reticulum (ER) where the A domain enters the cytosol to modify a cytosolic host substrate, leading to host cell intoxication or apoptosis29 (Figure 2).
Figure 2. Endocytosis of AB protein toxins.
AB5 protein toxins (Ctx, SubAB, Stx, and Ptx) enter by clathrin-mediated endocytosis (CME) and clathrin-independent lipid-raft (LR) pathways. Stx enters endosomal tubules (TB) and retrograde traffics with GPP130 through a bypass pathway towards the Golgi apparatus (GA). Other AB5 protein toxins enter a retrograde trafficking pathway towards the endoplasmic reticulum (ER). The A domain of AB5 protein toxins, excluding SubAB, retrograde translocate into the cytosol to modify a host substrate. SubAB modifies a resident ER protein. Atx associates with the plasma membrane (PM) phospholipids or lipid raft glycosphingolipids. Rtx and clostridial neurotoxins (CNT) enter early endosomes (EE) (in neurons); Rtx can retrograde traffic towards the ER or recycle back to the cell surface. CNTs escape acidifying endosomes (synaptic vesicles for BoNTs and endosomes in inhibitory neurons for TeNT) into the cytosol to cleave SNARE proteins. Other protein toxins (not shown) enter degradative pathways in late endosomes (LE).
Other AB toxins enter a retrograde endosomal pathway and traffic to the Golgi and the ER. For AB5 toxin family members, Ctx/LTx, Stx, Ptx, and SubAB, the B pentamer encodes a binding domain, but lacks a translocation domain2,29–32. Once these AB5 toxins retrograde traffic to the ER, the A domain dissociates from the B pentamer and interacts with a chaperone such as protein disulfide isomerase (PDI) to reduce the intrachain disulfide33–37. Following dissociation and release, the A1 subdomain usurps the Unfolded Protein Response (UPR) pathway to escape through the Sec61 translocon38. Upon retrograde translocation into the cytosol, via the Sec61 retrieval system, the A1 subdomain binds cytosolic chaperones, including Arf39,40, and refolds before trafficking to ADP-ribosylate the stimulatory heterotrimeric G-protein alpha subunit (Gαs). ADP-ribosylation locks Gαs in the GTP bound form to constitutively activate host adenylate cyclase and stimulate cAMP, uncoupling host signaling40. An exception within the AB5 family is SubAB which remains in the ER lumen to modify a chaperone protein involved in UPR41,42. The Ctx family members contain a KDEL, RDEL, or HDEL sequence at the C terminus of the A domain, which has been associated with retention in the ER by the KDEL receptor. The B pentamer also independently retrograde traffics to the ER, which indicates the KDEL sequence(s) optimize retrograde trafficking, but are not necessary nor sufficient to deliver AB5 toxins to the ER43,44.
Glycolipids
Eukaryotic cells are enveloped by a carbohydrate-rich coating, the glycocalyx. The glycocalyx contains glycoproteins and glycosphingolipids anchored in the outer leaflet of the plasma membrane45. Glycans face the exterior milieu and mediate cell-cell interactions via binding complementary molecules on other membranes or regulating activity by forming a buffer zone between host cell and extracellular microbes. Pathogens utilize a capsular glycocalyx coating to avoid host detection and immune cell assault and have evolved to utilize the host cell glycocalyx as receptors, by binding these molecules with adhesins, hemagglutinins, or by releasing protein toxins46.
Although glycosphingolipids refer to sphingosine-based lipids containing glycosidic bonds to simple carbohydrate molecules, the lipids are further classified into series based on a core structure47. The lipid portion of the molecule, ceramide, contains a sphingoid base connected to a fatty acid through an amide bond48. The ceramide portion of the ganglioside can be composed of sphingonine, sphingosine, or phytosphingosine and fatty acids can vary in chain length from ~16->24 carbons, which is associated with differences in trafficking48,49. Glycosphingolipids are abundantly found in the grey matter of the brain, with 10–12% of brain sphingolipids being composed of gangliosides50. Ganglio-series gangliosides have a common core of Galβ1–3GalNAcβ1–4Galβ1–4GlcβCer, while another common glycolipid species, the globosides, have Galα1–4Galβ1–4GlcβCer (Gb3) as a core45 (Figure 3).
Figure 3. Simple- and complex- gangliosides.
Gangliosides are glycosphingolipids which contain a ceramide and oligosaccharides that are complexed with one or more sialic acid (ceramide not depicted). Simple gangliosides are depicted on the top row and more complex gangliosides are depicted on bottom row. Gangliosides are enriched in neurons.
While gangliosides are not required for cell survival, gangliosides are necessary for survival in whole-organisms51. The hydrophilic, negatively charged headgroups of gangliosides spread on the cell surface and exist in varied conformations at each branch points. Gangliosides are negatively charged due to decoration with 9-carbon sugars, sialic acids, off the core sugars50. In humans, sialic acids are composed of N-acetylneuraminic acid (Neu5Ac) while other mammals also synthesize N-glycolylneuraminic acid (Neu5Gc)52. In addition to variation in ceramide composition, the number of sialic acid groups connected to the core sugars also vary. Gangliosides contain 75% of the brain’s sialic acid, primarily GM1a, GD1a, GD1b, and GT1b (Figure 3)53. Gangliosides like other glycosphingolipids are restricted to the outer leaflet of the cell membrane, with the ceramide and first core sugar glucose engaged in this lipid layer51. Gangliosides are primarily found in the extracellular plasma membrane, but are also detected in intracellular organelles. The Golgi is responsible for the diversity of cell surface glycans54. Neither gangliosides nor globosides are uniformly distributed in the plasma membrane, but cluster in lateral microdomains called detergent-insoluble glycosphingolipid-enriched domains (DIGs) which self-associate due to biophysical properties of long, unsaturated carbon chains55. DIGs are known as lipid rafts in living cells and concentrate glycosylphosphatidylinositol-(GPI) anchored proteins, cholesterol, and gangliosides56. Lipid rafts are often coupled to signaling complexes, which contain receptor tyrosine kinases (RTKs), G-protein coupled receptors (GPCRs), and downstream partners such as mitogen-activated protein (MAP) kinases and heterotrimeric G-proteins (GαPγ). The binding of protein toxins to glycosphingolipids is prerequisite, but coupling of the lipid species to signaling domains for productive uptake affects tissue susceptibility.
Complex gangliosides, GD1a and GT1b, are ligands for myelin-associated glycoprotein (MAG), promoting axon stability and neurite inhibition57. Altered surface expression or storage of gangliosides is associated with abnormal cell-cell contact inhibition and various diseases58. Mice lacking some complex gangliosides, GM2 or GD2 synthase knockouts, develop hearing loss and abnormal motor pathology with age59. Double knockouts in these aforementioned glycan synthase systems (St3gal5 or St8Sia1/B4galnt1) die shortly after birth; therefore, the necessity for ganglioside synthesis is commonly exploited by pathogens and toxins that bind carbohydrates60–62.
Protein Toxins utilize Gangliosides as Host Receptors
Clostridial Neurotoxins of Clostridum botulinum and C. tetani The clostridial neurotoxins (CNTs) of Clostridum botulinum and tetani target motor neuron presynaptic termini in the periphery and inhibitory interneurons in the central nervous system, respectively8. CNTs comprise seven serologically distinct BoNTs (A-G) and TeNT63. CNTs are synthesized as single-chain AB toxins cleaved into a di-chain; an A catalytic light chain (LC) connected to a B heavy chain (HC) containing an N-terminal translocation domain and a C-terminal receptor binding domain64. The LCs of CNTs are zinc-metalloproteases, which cleave neuron-specific isoforms of SNARE proteins involved in vesicle fusion at the plasma membrane65,66. Following cleavage, the tertiary complex cannot form between vesicle and plasma membrane-associated SNARES, and fusion of neurotransmitter-containing vesicles to the presynaptic membrane is inhibited, resulting in paralysis67,68. Toxin trafficking determines the type of paralysis, as both BoNT/B and TeNT cleave the host SNARE at the same scissile bond66,69. Flaccid paralysis is characterized by inhibition of acetylcholine release at the neuromuscular junction, with the motorneuron not signaling the downstream muscle to contract70. For TeNT, signaling is uncoupled between an interneuron and motorneuron when release of inhibitory glycine into the synaptic cleft is inhibited. This results in spastic paralysis as the motor neuron is persistently releasing acetylcholine, causing muscle contraction70. CNTs bind complex gangliosides, which are enriched in the outer leaflet of neurons71.
TeNT, when isolated from brain and cultured neuronal cell membranes, was found bound to a glycolipid, later identified as GT1b72,73. TeNT has two ganglioside binding pockets in the receptor binding domain74. Toxin binding two gangliosides is necessary and sufficient for entry into neuronal and non-neuronal cells20. Although protein receptors for the toxin have been proposed, none are necessary for intoxication75–77. TeNT has been observed to colocalize with signaling endosome proteins such as the growth factor receptors, which like gangliosides, are enriched on lipid rafts and DIGs78. Recent studies indicate that the receptor binding domain pivots with respect to the N-terminal domains of full-length TeNT in response to a change in pH79. This may influence the orientation of the translocation domain with respect to channel formation, as BoNT-receptor interactions regulate channel formation80,81. Botulinum neurotoxins, excluding BoNT/C, bind a ganglioside and a protein receptor sequentially; upon depolarization of the neuronal membrane, fusion of synaptic vesicle exposes the lumenal domains of a synaptic vesicle protein receptors82. Interactions with both the ganglioside and SV2 appear to be sequential for entry to neurons83. The overall structure of BoNT/C aligns well with other BoNTs serotypes, except at the C terminus where there are two unique ganglioside binding pockets, which appear sufficient for the entry of BoNT/C into neurons84, precluding the use of a host protein receptor.
Cholera toxin family of AB5 toxins (Ctx/LTx) of V. cholerae and enterotoxigenic Escherichia coli The cholera toxin family is comprised of Ctx produced by Vibrio cholerae and the related heat-labile enterotoxins (LTx), LT-I, LT-IIa, LT-IIb, and LTIIc, produced by enterotoxigenic Escherichia coli (ETEC)2. V. cholerae and ETEC are often found in water and food contaminated with fecal matter. Ctx/LTx target intestinal epithelial cells to cause diarrheal disease, with LTx family members generally eliciting pathology of shorter duration and milder illness than Ctx85. Ctx and LTx cause distention in the rabbit illeal loop model, previously a gold standard for diagnosing this disease86. Cholera outbreaks are endemic in many developing countries, while LT-related illnesses are associated with travelers’ diarrhea in developing countries and more recently, livestock disease in developed countries86.
Ctx and LT-I invade the intestinal epithelia by binding with high affinity to the ganglioside GM1a and retrograde traffic to the ER21,87. Ctx/LTx A1 catalytic domain is an ADP-ribosyltransferase that ADP-ribosylates Arg177 of the stimulatory α-subunit of the heterotrimeric G-protein, Gαs, following translocation into the cytosol88–91. Addition of the negatively charged ADP-ribosyl group constitutively activates Gαs by preventing hydrolysis of GTP. Gαs-GTP binds to host adenylate cyclase, which increases cyclic AMP (cAMP) production by the cyclase and modulates ion channels to release ions into the extracellular space92. Persistent chloride ion release and efflux of water out of the intestinal epithelial cells and into the intestine results diarrhea, electrolyte imbalances, and severe dehydration93.
Ctx- and LT-I- B subunits share ~ 80% identity and both bind GM1a; however, LT-I also binds LPS and A/B glycolipid antigens on human erythrocytes21,94–97. The B subunits are less conserved between Ctx and LT-II family members with ~ 14% shared identity. LT-IIa predominantly binds GD1b, GD1a and GT1b, and to a lesser extent GM1a and other gangliosides, while LT-IIb predominantly binds only GD1a and GD1b22,98. The newest member LT-IIc has been observed to bind GM1 species with long acyl chains99. Ctx/LTx bind up to 5 glycans subunits; however, for CTx, binding of one GM1a molecule is sufficient, though less efficient for retrograde trafficking and cell intoxication43.
Stx of Shigella dysenteriae and Stx1/2 of Shiga toxin-producing E. coli Stx, produced by Shigella dysenteriae, was identified by Kiyoshi Shiga as the causative agent of dysentery100. Almost a century later, the toxins now renamed as Shiga-like toxins 1 and 2 (Stx1 and Stx2), were identified in Shiga toxin-producing E. coli (STEC). Stx and Stx1 differ by a single amino acid, while Stx1 and Stx2 share ~55% identity30. Both E. coli Stx1 and Stx2 produce immunologically distinct subtypes. Stx-producing bacteria cause the infection Shigella, which is contracted by ingestion of contaminated foods101.
Stxs are AB5 toxins, the A subunit contains N-glycosidase activity responsible for removal of an adenine residue from 28S rRNA of the 60S ribosome subunit102. Modification of the rRNA inhibits protein synthesis as the ribosome no longer interacts with elongation factor 1103,104. Stalled synthesis induces ribotoxin stress and the UPR, resulting in apoptosis105. Stx translocates through intestinal epithelial cells to target the endothelial cell layer, and in serious cases intoxicates the kidneys101. Following modification of the rRNA substrate, the localized infection presents as hemorrhagic diarrhea and if Stx reaches the kidneys, hemolytic uremic syndrome (HUS) occurs as a result of low platelet count, erythrocyte lysis, and kidney failure100. The five identical B pentamers of Stx bind the globoside Gb3, Stx1 and Stx2 weakly bind Gb4 (GalNAcβ1–3Galα1–4Galβ1–4GlcβCer), while subtype Stx2e preferentially binds globoside Gb4106. Cells and mice lacking Gb3 are insensitive to Stx, while mice with Fabry’s disease containing excess Gb3 throughout tissues also possess increased resistance to Stx107. Gb3, like other glycosphingolipids, clusters in DIGs or lipid rafts in cells108. The presence of cholesterol in microdomains enhances binding and entry of the Stx and Stx1, as pre-treatment with cholesterol-binding agents such as methyipcyclodextrin (MβCD) reduces Golgi-localization and cytotoxicity109. Stx2 is associated with lipid rafts in vitro and clinically associated with development of HUS; Stx2 does not interact with the host trans-Golgi cycling resident protein, GPP130106,110,111
Shiga toxins are endocytosed by clathrin-dependent and clathrin-independent pathways, depending on the cell type32,112. In HeLa cells, clathrin-independent tubular invaginations contribute to sorting of Stx B subunits, which bind GPP130 to facilitate retrograde traffic113,114.
Ricin of Ricinus communis Ricin (Rtx) is a toxigenic lectin produced by seeds of the castor oil plant Ricinus communis. Exposure to Rtx by inhalation or injection is more lethal than digestion, as the toxin is sensitive to bile acids115. The ricin A subunit, like Stx, is a type II ribosome-inactivating protein, which inhibits protein synthesis by depurination of the rRNA catalyzed by N-glycosidase activity102,116. Stx- and Rtx- A subunit share < 50% identity, but residues involved in adenine abstraction are conserved103. Cytotoxic effects of include cell lysis through blebbing and DNA degradation27. The route of exposure determines the symptoms; intravenous injection in rodents results in hypoglycemia, inflammation, and ketosis, while inhalation results in fluid accumulation in the lungs and eventual respiratory failure117. Like other single-chain AB toxins, Rtx A subunit catalytic domain connects to the B subunit binding domain by a peptide linker. The B domain contains two homologous lectin-binding sites specific for galactose117–119. Rtx undergoes post-translational glycosylation on both chains, which remain associated following host cleavage by an interchain disulfide bridge4,23.
Rtx binds glycolipids (including gangliosides) and glycoproteins containing terminal galactose (or N-acetylgalactosamine) in βi-4-linkage. In early endosomal sorting, Rtx enters a recycling pathway and traffics back to the cell surface as well as to lysosomes120. Uptake of ricin occurs by two pathways: ricin B subunit binds galactose-containing host glycolipid and glycoproteins and conversely, glycosylated residues of ricin toxin are bound by a host mannose receptor, which cycles between trans-Golgi and late endosomes121,122. While galactose-binding by the B subunit results in robust uptake of Rtx, the latter mannose-dependent uptake by the mannose receptor is responsible for the cytotoxic effects by delivering RTx to a productive trans-Golgi pathway121.
SubAB of STEC Recently, a new AB5 toxin member, subtilase cytotoxin (SubAB) was found to cause HUS-like illness in STEC strain O113:H21. SubAB binds non-human Neu5Gc as a receptor and is more cytotoxic than Stx in in vitro cell culture assays2. SubAB is also unique, as the A subunit is not required to translocate into the cytosol, since the target substrate is located within the lumen of the ER. The A subunit contains subtilase-like serine protease activity and cleaves the ER-resident protein BiP42. BiP contributes to the UPR as a chaperone for folding nascent proteins to be degraded by the proteasome123. BiP also forms a seal within the Sec61 pore; proteolysis of BiP destroys the integrity of this barrier leading to ER permeability123. Humans, in contrast to the great apes, have acquired mutations in the biosynthetic pathways to synthesize and convert Neu5Ac into Neu5Gc52. With loss of Neu5Gc, Neu5Ac becomes the only sialic acid synthesized. Intriguingly, humans are susceptible to SubAB due to incorporation of dietary Neu5Gc into cellular glycolipids and proteins via meat and dairy consumption124.
While SubAB specifically recognizes the sialic acid Neu5Gc, which contains an additional hydroxyl substituent relative to Neu5Ac, many human pathogens have evolved to exploit Neu5Ac as a receptor. For example, the newest member of the enterotoxins, LT-IIc was isolated from avians, some of which produce Neu5Gc, but has similar affinity to both sialic acid species99,125. Incorporation of dietary sialic acids and pathogens that bind multiple sialic acid species has potential to expand the host range for zoonotic diseases.
Ptx of Bordetella pertussis Ptx is a major virulence factor produced by B. pertussis, which is transmitted by inhalation of aerosolized droplets transmitted early in the infectious state of an infected individual126. Although Ptx alone does not cause whooping cough, entry into pulmonary epithelial and endothelial cells leads to inflammation and white blood cell recruitment127. Ptx alteration of host signaling also leads to hyperinsulinemia and pulmonary hypertension127,128.
Ptx is an AB5 family member. Like Ctx, the A subunit known as S1, ADP-ribosylates the inhibitory heterotrimeric G-protein (Gαi) at a C-terminal cysteine, causing dysregulation of cAMP signaling and ion efflux129. The B pentamer is comprised of four heterologous subunits known as S2, S3, S4, and S5, in a 1:1:2:1 ratio, respectively130. Heterogeneity of the B pentamer subunits, including the two carbohydrate binding subunits, S2 and S3, allows for Ptx to recognize multiple glycans with varied affinities131,132. As a result, Ptx associates with glycolipids and glycoproteins such as the serum protein fetuin, which is decorated with N-linked glycans24. B pentamer binding to host receptors alone can induce T cell proliferation and agglutinate erythrocytes133. S2 and S3 are 77% identical and each contain a sialic acid (Sia) binding site and an N-terminal carbohydrate binding site130. In a glycan array screen, Ptx bound α2–6 linked Sia in N-glycans, α2–3 linked Sia in O-glycans, and complex gangliosides134,135. To discriminate glycans bound by the N-terminal region, dimers lacking the Sia domain were assembled with S4 for stability and bound asialo N-linked glycans134.
Although not unique to Ptx, characterization of glycans bound by the B subunits of toxins varies depending on the format of the assay. For example, Ptx binds gangliosides, such as GD1a on liposomes, but unlike Ctx, does not detectably bind lysate-derived gangliosides separated by thin layer chromatography (TLC)135,136. Additional considerations for interpretation of toxin affinity for glycan substrates are the orientation of the glycan, flat or curved surfaces, as well as the linker composition, ceramide, or acyl species of the glycolipids.
Alpha-toxin of C. perfringens C. perfringens produces many virulence factors including alpha-toxin, a virulence factor for pathogenesis. Alpha-toxin is the causative agent of gas gangrene, which is characterized by edema, myonecrosis, and gas. Inflammation, resulting from alpha-toxin action leads to septic shock and organ failure137,138. Alpha-toxin is also associated with Crohn’s Disease, where the toxin targets intestinal epithelial cells and resident macrophages139.
Alpha-toxin is a single-chain AB toxin, which contains an A subunit with zinc-metalloprotease-dependent phospholipase C catalytic activity, which hydrolyzes phospholipids into diacylglycerol (DAG) and ceramide. The A subunit is connected to a binding domain that contains both hemolytic activity and a C2-like polycystin-1, lipoxygenase, alpha-toxin (PLAT) domain137. In the presence of calcium, the B domain binds and partially inserts into the phospholipid membrane140. There is also a ganglioside binding site, specific for GM1a, located on a loop between the A and B subunits141. The generation of DAG, a common eukaryotic second messenger, results in uncoupled signaling and release of cytokines142.
Molecular interactions of Protein toxins with sugars
The AB5 family members Ctx and SubAB contain five sugar binding sites, one per B subunit, while Ptx contains up to two functional binding sites on each of the S2 and S3 subunits, and Stx contains up to 3 binding sites per subunit106,130,143,144. Each AB5 family member B subunit contains a homologous fold at their C terminus comprised of five or six antiparallel beta strands, which form a barrel topped by a short alpha helix (Figure 4, panel A) with and without sugar. This oligonucleotide and oligosaccharide binding domain is known as Oligomer Binding Fold (OBF)145. Ptx S2 and S3 subunits contain an additional N-terminal domain with homology towards C-type eukaryotic lectins such as rat mannose binding protein134.
Figure 4. AB5 toxin-aromatic sugar interactions.
Crystal structures of the B pentamers of AB5 toxins LT-I, Ctx, SubAB, Ptx, and Stx1 are shown in blue on the left of each panel. Sugar ligands (carbohydrates, gangliosides, or globosides) are shown in yellow. Aromatic residues in the B domain are shown in green and numbered on the right. Polar groups, hydrophobic interactions, and inter-domain interactions are left out for clarity. Selected oxygen atoms are shown in red and nitrogen atoms shown in blue. A) LT-1 B pentamer (PDB: LTT1) contains an oligosaccharide binding fold (OBF) shown in green and is bound to lactose. The oligosaccharide binding site is shown magnified on the right. B) Ctx B pentamer (PDB: 3CHB) shown in blue with one subunit bound to GM1a. A single subunit is shown on right with aromatic residues involved in binding. C) SubAB B pentamer (PDB: 3DWP) is bound to Neu5Gc, with magnification of this interaction shown on right. D) Ptx B pentamer (PDB: 1PTO) is shown with S3 bound to GM1a. S2 and S3 interactions with GM1a are similar; S3-GM1a is shown on the right. E) Stx1 B pentamer (PDB: 1BOS) is shown bound to three Gb3 molecules. One B subunit interacts with three Gb3 molecules shown on the right.
Cholera toxin family of AB5 toxins (Ctx/LTx) of V. cholerae Ctx and LT utilize a pocket formed by loops to bind GM1a; in Ctx these loops are found between β1 to α1, β4 to α2, β5 to β6 on one monomer and the loop between β2 to β3 on the other monomer144. The pocket is oriented towards the glycolipid receptor at the membrane. For Ctx, the terminal sugars of GM1a, galactose and sialic acid, predominantly interact with the B subunit2,146. The galactose forms hydrogen bonds via oxygen with the nitrogen atoms in the B subunit residues Asn90 and Lys91143. A hydrophobic interaction between galactose and the indole ring of Trp88 is also observed (Figure 4, panel B). The sialic acid of GM1a forms hydrogen bonds with the carbon backbone of residues Glu11 and His13, and a hydrophobic interaction with Tyr1294. Gly33 lies in a loop between β2 and β3 of the adjacent monomer and forms the pocket edge with one hydrogen bond to galactose94,146. In addition to GM1a, LT binds asialo-GM1 and GD1b, derivatives that lack and gain an additional sialic acid, respectively; the substitution of His13 for Arg13 in LT is suspected to accommodate this promiscuity147. LPS binding and erythrocyte group A or B antigens by LT have been mapped to residues not involved in binding GM1a148.
SubAB of STEC Unlike Ctx and Stx, SubAB binds the non-human derived sialic Neu5Gc in a shallow binding pocket located on outer face of the B pentamer41. The SubAB pocket resembles the sialic acid pocket of Ptx B subunits S2, S3, and S5, where S2 and S3 bind sialic acid Neu5Ac in a similar orientation, but unique interactions are involved in binding the different sialic species41. The chemical structure of Neu5Gc differs from human-derived Neu5Ac by a single hydroxyl group52. However, the presence of this hydroxyl group allows SubAB pocket residues Tyr78 and Met10 to form hydrogen bonds with sialic acid, where Neu5Gc stacks with Phe11 of SubAB, forming a van der Waals interaction (Figure 4, panel C)2,41. The affinity of SubAB for Neu5Gc is unique and specific; serum from chimpanzees containing glycolipids and proteins decorated with Neu5Gc, but not human serum that predominantly contains Neu5Ac, can compete with SubAB binding of immobilized the Neu5Gc glycan124.
Ptx of Bordetella pertussis Ptx binds glycoproteins through B subunits, S2 and S3, which each contain two unique carbohydrate domains. Analysis of a crystal structure incubated with glycolipids identified electron density for Neu5Aca2–6Gal, but not galactose within the binding pocket130,131 (Figure 4, panel D). Sialic acid forms hydrogen bonds with Ser104, and hydrophobic interactions with Tyr102 and Tyr103, on the sides of the pentamer130,131. It was previously observed that deletion of Tyr102 and Tyr103 in S2 and S3, or substitution of Ser104 with a charged group reduced biological activity of Ptx149.
Stx of Shigella dysenteriae Stx1 binds the Gb3 analogue in up to three binding sites per B subunit. Gb3 binding pockets are located near the bottom of the B pentamer to allow Stx1 to bind the membrane face150. Two binding sites are formed between subunits, while one is formed near the membrane. The first site is a pocket formed on a single monomer by loops β2–β3, β5–β6, and strands β3 and β4; hydrogen bonding is observed between both galactose sugars and the B subunit residues Thr21, Glu28, Gly60, and Asp17150. Additionally, hydrophobic interactions, including a ring stacking interaction with Phe30 are observed with Gb3106,150. The second Gb3 binding site is located between monomers and is comprised of loops β2-β3 from site one, β4-α, and β5-β6 (Figure 4, panel E)106,109. This site was originally predicted based on known crystal structures of other AB5 family members and appears to be highly occupied. The binding pocket contacts Gb3 through hydrogen bonding the two galactose sugars with the Asp16, Asn32, Arg33, Asn55, and Phe63 while many hydrophobic interactions are observed between the galactose sugars, glucose interacts with Asn55150. At binding site 3, Gb3 is bound within the pentamer cleft by β4-α of the second site, and β2-β3 loop at the first site. Asp18, Trp34 and Asn35 hydrogen bond with the first galactose while the residue Trp34 located within and on an adjacent monomer stack with both the first and second galactose moiety106,150.
Rtx of Ricinus communis The Rtx crystal structure exhibits two non-cooperative lectin binding domains containing a p-trefoil fold separated by ~3.5nm119. Rtx, binds glycolipids or glycoproteins, and the distance between lectin binding domains allows the B subunit to sterically accommodate host receptors151. The two domains are composed of three lobes referred to as 1(α, β, and γ) and 2(α, β, and γ), with 1α and 2γ confirmed to bind galactose23,118,152,153. Each lobe contains a tripeptide kink and an aromatic residue; the motif Gln,X,Trp118. For ricin, mutagenesis, docking analysis, and comparison to other lectin-binding proteins predicts two interactions upon receptor binding: ring stacking of the galactose moiety against an aromatic residue and hydrogen bonding between pocket residues and oxygen groups on the sugar151.
Clostridial Neurotoxins of Clostridum botulinum and C. tetani Alignment of the clostridial toxins alpha-toxin, TeNT, and most BoNTs serotypes, reveals a conserved GM1a binding motif comprised of ~30 residues (H…SXWY…G)19,141. TeNT and BoNTs contain this motif within the receptor binding domain, while alpha-toxin contains this sequence on an exposed loop between the A and B subunits64,137. For all toxins, mutagenesis of the aromatic consensus residues tryptophan and tyrosine abrogates binding to the glycolipid receptor.
Tetanus toxin binds the membrane utilizing two pockets. Mutation of Trp1289Ala (W pocket) and Arg1226Leu (R pocket) abrogates ganglioside binding in solid phase and cell binding assays20,74. Mutational analysis of single pockets in TeNT discriminated glycan preference bound by each pocket; the R pocket binding b-series and W pocket binding a-series gangliosides74. The crystal structure reveals that the R pocket binds a GT1b analogue in a shallow pocket primarily through the sialic acid moiety. The bound sialic acid forms hydrogen bonds through oxygens to charged and polar residues Asp1147, Asp1214, Asn1216, and Tyr122974. A salt bridge is also formed between the sialic acid and Arg122674,154. The W pocket, which is also conserved in many BoNTs binds a lactose moiety by stacking with the indole ring of Trp1289 and forming hydrogen bonds between the sugar’s hydroxyl group and the charged side chains of His1272, Asp1222 and the amide of Thr127074,154. The recently solved crystal structure of full length TeNT includes GD1a within the W pocket which stabilized the crystal79. A similar interaction is observed in the W pocket of BoNT/A incubated with a GT1b analogue; analogous hydrogen bonding to His1253, Glu1203, and Ser1264 and hydrophobic interactions with the indole ring of Trp1266155.
Alpha-toxin does not bind to sialic acids alone or require complex gangliosides, but is specific to GM1a141. Molecular docking simulation predicts that alpha-toxin and GM1a interaction is dependent upon hydrophobic ring stacking and hydrogen bonding of Trp84 in alpha toxin with the sialic acid, while Tyr85 hydrogen bonds with the galactosamine portion of the glycolipid156. The binding of GM1a is not necessary for endocytosis of alpha-toxin, but is hypothesized to tether the toxin, before binding to phospholipids via C-terminal aromatic residues140. Binding of alpha toxin to GM1a is necessary to induce signaling through TrkA, a major step in pathogenesis142.
Protein Toxin-mediated trafficking and signaling
Gangliosides located within lipid rafts are coupled with signaling complexes, which include RTKs and GPCRs that modulate downstream signaling upon ligand binding. Interaction between ganglioside and receptor modulates activity through glycan-glycan or glycan-protein interactions; such interactions include hydrogen bonding that interferes with normal ligand-receptor interactions157,158. For example, how GM3 modulates fibroblast growth factor receptor or epidermal growth factor receptor may be due to reducing receptor-ligand binding affinity or by exclusion of receptors from lipid rafts that leads to dampened signal159,160. Conversely, the ganglioside GM1a stimulates neurotrophin receptors such TrkA like the endogenous ligand, nerve growth factor (NGF), and leads to receptor activation161. The studies that elucidated gangliosides as ligands for TrkA found brain slices stimulated with high concentrations of whole-ganglioside, but not the components such as sialic acids, glycans, or ceramides, induced autophosphorylation of TrkA leading to downstream signaling through MAP kinase pathway161. In addition to being ligands for growth factor receptors, gangliosides are also ligands for MAG, inhibiting neuritogenesis and other differentiation processes57.
Some protein toxins signal through gangliosides at the membrane to recruit members of endocytic pathways. TeNT receptor binding domain colocalizes with growth factor receptors p75NTR and TrkA in a well-characterized neurotrophic retrograde pathway toward the soma in peripheral motor neurons78. TeNT receptor binding domain also stimulates activation of the TrKA receptor in cultured primary neurons162. Upon addition of TeNT, dose-dependent TrkA autophosphorylation and activation of phospholipase C leads to activation of MAP kinase ERK1/2 and other neuroprotective pathways163,164. TeNT activation of TrkA may mimic the binding of NGF and induce uptake into signaling endosomes toward the retrograde traffic.
Alpha-toxin is associated with lipid rafts. The binding of phospholipids by a C-terminal domain and hydrolysis causes negative membrane curvature and induces endocytosis138. Alpha-toxin also binds ganglioside GM1a through a conserved ganglioside motif. This secondary binding of GM1a activates TrkA, causing a phosphorylation cascade of MAP kinase substrates142. Unlike TeNT, alpha-toxin signaling is associated with release of inflammatory chemokines and reactive oxygen species, which is responsible for the severity of gas gangrene symptoms. This signaling is mediated through GM1a, as treatment of cells with neuramidase to cleave the terminal sialic acid or an inhibitor of ganglioside synthesis, PPMP, dose-dependently reduces cytokine release upon exposure to alpha toxin141. Mutation of Trp84-Tyr85 attenuates GM1a binding, activation of TrkA signaling, and release of IL-8141. Once alpha-toxin binds the membrane, the phospholipase activity forms DAG which leads to alterations in the membrane that may induce vasculature permeability leading to edema138.
AB5 Toxins and Ganglioside Cross-linking
Crosslinking gangliosides can induce endocytosis through receptor activation and can enrich lipid rafts for TrkA165. Ctx clusters ganglioside GM1a by binding five molecules concurrently, leading to more efficient receptor-mediated internalization, and in neurons, the activation of TrkA43,166. Stx binds up to 15 molecules of Gb3 at the plasma membrane, which activates the RTK Syk, which phosphorylates clathrin heavy chain150,167. This leads to recruitment of adaptor protein 2 and clathrin-coated vesicles to the membrane, and may be responsible for the clathrin-dependent entry of Stx in some cell lines168. In some cases, fatty acyl chain saturation and chain length are as important as the carbohydrate moiety for toxin binding, entry, and trafficking into productive pathways. One study utilized lipid probes of varying chemical species and acyl lengths to demonstrate that the ceramide portion of GM1a alters lipid trafficking within the cell49; unsaturated ceramide chains retrograde trafficked from plasma membrane to the ER more efficiently than long, saturated ceramides, which were not localized with raft-associated proteins49. In contrast, Ctx B pentamer did not alter lipid localization, which may explain how toxins exploit lipids trafficking to different compartments.
AB5 Toxins and Lipid trafficking
Lipids properties directly affect raft composition, which affect ganglioside dynamics by retaining certain lipids and GPI-anchored protein species169. Ceramide composition and fatty acid acyl length also affects partitioning properties, cargo sorting, and further increase heterogeneity of the plasma membrane49. For Stx1 and Stx2, in vitro binding affinity of Gb3 ganglioside was increased for shorter acyl chains compared to longer acyl chains170,171. Gangliosides with shorter acyl chains partition differently within a cell membrane, leading to alterations in ligand accessibility172. Chain length may play a physiological role in the association of Stx with lipid rafts in cells, which are also required for clathrin-independent Stx uptake167,168. The disruption of Gb3 localization to lipid rafts protects cells against Stx intoxication, but does not affect binding of the toxin173. LT-IIb binds non-raft associated GD1a in cultured T84 human colon carcinoma cells; LT-IIb binds, but does not enter a productive retrograde pathway unlike Ctx which binds GM1a174. These studies may resolve how host glycolipids as ubiquitous as gangliosides serve to bind and target protein toxins to certain tissues, causing the hallmarks of disease.
Table 1:
AB Protein Toxin Receptors, Action, and Disease
| Organism | Toxin | Receptor(s) | Substrate | Enzyme action | Disease |
|---|---|---|---|---|---|
| C. tetani | TeNT | GT1b | VAMP | Zn2+ metalloprotease |
spastic paralysis |
| C. botulinum | BoNT/A | GT1b, SV2C | SNAP-25 | Zn2+ Metalloprotease |
flaccid paralysis |
| C. perfringens | Atx | GM1a | Phosphatidyl choline | Phospholipase C | gas gangrene |
| V. cholerae | Ctx | GM1a | Gαs | ADP-ribosyl transferase | diarrheal disease |
| ETEC | LT-I LT-IIa LT-IIb LT-IIc |
GM1a, LPS, A/B blood GD1b>GD1a>GM1a GD1a GM1a, long acyl |
Gαs | ADP-ribosyl transferase | diarrheal disease |
| B. pertussis | Ptx | glycolipids, glycoproteins; galactose, lactose, GD1a | Gαi | ADP-ribosyl transferase | whooping cough |
| R. communis | Rtx | glycolipids, glycoproteins; galactose | A4256 28S rRNA |
RNA N-glycosidase | pulmonary edema, seizures |
| S. dysentariae | Stx | Gb3 | A4256 28S rRNA |
RNA N-glycosidase | hemorrhagic colitis, HUS |
| STEC | Stx1 Stx2 |
Gb3 Gb3, Gb4 |
A4256 28S rRNA |
RNA N-glycosidase | hemorrhagic colitis, HUS |
| STEC | SubAB | glycolipids, glycoproteins; Neu5Gc | BiP | Subtilase-like serine protease | HUS |
References
- 1.Smyth CJ. The Comprehensive Sourcebook of Bacterial Protein Toxins Edited by Alouf Joseph E. and Popoff Michael R. San Diego: Academic Press, 2006. 1047pp., illustrated. $199.95 (hardcover). Clinical Infectious Diseases. 2006;43(5):669–669. [Google Scholar]
- 2.Beddoe T, Paton AW, Le Nours J, Rossjohn J, Paton JC. Structure, biological functions and applications of the AB5 toxins. Trends Biochem Sci. 2010;35(7):411–418. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Garred ∅, Dubinina E, Polesskaya A, Olsnes S, Kozlov J, Sandvig K Role of the Disulfide Bond in Shiga Toxin A-chain for Toxin Entry into Cells. Journal of Biological Chemistry. 1997;272(17):11414–11419. [DOI] [PubMed] [Google Scholar]
- 4.Lappi DA, Kapmeyer W, Beglau JM, Kaplan NO. The disulfide bond connecting the chains of ricin. Proceedings of the National Academy of Sciences of the United States of America. 1978;75(3):1096–1100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Schiavo G, Papini E, Genna G, Montecucco C. An intact interchain disulfide bond is required for the neurotoxicity of tetanus toxin. Infection and Immunity. 1990;58(12):4136–4141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Tomasi M, Battistini A, Araco A, Roda LG, D’Agnolo G. The Role of the Reactive Disulfide Bond in the Interaction of Cholera-Toxin Functional Regions. European Journal of Biochemistry. 1979;93(3):621–627. [DOI] [PubMed] [Google Scholar]
- 7.Zuverink M, Chen C, Przedpelski A, Blum FC, Barbieri JT. A Heterologous Reporter Defines the Role of the Tetanus Toxin Interchain Disulfide in Light-Chain Translocation. Infect Immun. 2015;83(7):2714–2724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Montecucco C, Schiavo G. Mechanism of action of tetanus and botulinum neurotoxins. Molecular Microbiology. 1994;13(1):1–8. [DOI] [PubMed] [Google Scholar]
- 9.Klimpel KR, Molloy SS, Thomas G, Leppla SH. Anthrax toxin protective antigen is activated by a cell surface protease with the sequence specificity and catalytic properties of furin. Proceedings of the National Academy of Sciences of the United States of America. 1992;89(21):10277–10281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Gordon VM, Klimpel KR, Arora N, Henderson MA, Leppla SH. Proteolytic activation of bacterial toxins by eukaryotic cells is performed by furin and by additional cellular proteases. Infection and Immunity. 1995;63(1):82–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Gordon VM, Leppla SH. Proteolytic activation of bacterial toxins: role of bacterial and host cell proteases. Infection and Immunity. 1994;62(2):333–340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Lencer WI, Constable C, Moe S, et al. Proteolytic Activation of Cholera Toxin and Escherichia coli Labile Toxin by Entry into Host Epithelial Cells: SIGNAL TRANSDUCTION BY A PROTEASE-RESISTANT TOXIN VARIANT. Journal of Biological Chemistry. 1997;272(24):15562–15568. [DOI] [PubMed] [Google Scholar]
- 13.Lea N, Lord JM, Roberts LM. Proteolytic cleavage of the A subunit is essential for maximal cytotoxicity of Escherichia coli 0157:H7 Shiga-like toxin-1. Microbiology. 1999;145(5):999–1004. [DOI] [PubMed] [Google Scholar]
- 14.Burns DL, Manclark CR. Role of cysteine 41 of the A subunit of pertussis toxin. J Biol Chem. 1989;264(1):564–568. [PubMed] [Google Scholar]
- 15.Kozaki S, Kamata Y, Watarai S, Nishiki T-i, Mochida S. Ganglioside GT1b as a complementary receptor component forClostridium botulinumneurotoxins. Microbial Pathogenesis. 1998;25(2):91–99. [DOI] [PubMed] [Google Scholar]
- 16.Mahrhold S, Bergstrom T, Stern D, Dorner BG, Astot C, Rummel A. Only the complex N559-glycan in the synaptic vesicle glycoprotein 2C mediates high affinity binding to botulinum neurotoxin serotype A1. Biochem J. 2016;473(17):2645–2654. [DOI] [PubMed] [Google Scholar]
- 17.Mahrhold S, Rummel A, Bigalke H, Davletov B, Binz T. The synaptic vesicle protein 2C mediates the uptake of botulinum neurotoxin A into phrenic nerves. FEBS Lett. 2006;580(8):2011–2014. [DOI] [PubMed] [Google Scholar]
- 18.Peng L, Berntsson RPA, Tepp WH, et al. Botulinum neurotoxin D-C uses synaptotagmin I and II as receptors, and human synaptotagmin II is not an effective receptor for type B, D-C and G toxins. Journal of Cell Science. 2012;125(13):3233–3242. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Rummel A, Hafner K, Mahrhold S, et al. Botulinum neurotoxins C, E and F bind gangliosides via a conserved binding site prior to stimulation-dependent uptake with botulinum neurotoxin F utilising the three isoforms of SV2 as second receptor. J Neurochem. 2009;110(6):1942–1954. [DOI] [PubMed] [Google Scholar]
- 20.Chen C, Fu Z, Kim JJ, Barbieri JT, Baldwin MR. Gangliosides as high affinity receptors for tetanus neurotoxin. J Biol Chem. 2009;284(39):26569–26577. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Kuziemko GM, Stroh M, Stevens RC. Cholera toxin binding affinity and specificity for gangliosides determined by surface plasmon resonance. Biochemistry. 1996;35(20):6375–6384. [DOI] [PubMed] [Google Scholar]
- 22.Fukuta S, Magnani JL, Twiddy EM, Holmes RK, Ginsburg V. Comparison of the carbohydrate-binding specificities of cholera toxin and Escherichia coli heat-labile enterotoxins LTh-I, LT-IIa, and LT-IIb. Infect Immun. 1988;56(7):1748–1753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Frankel A, Tagge E, Chandler J, Burbage C, Willingham M. Double-site ricin B chain mutants retain galactose binding. Protein Eng. 1996;9(4):371–379. [DOI] [PubMed] [Google Scholar]
- 24.Witvliet MH, Burns DL, Brennan MJ, Poolman JT, Manclark CR. Binding of pertussis toxin to eucaryotic cells and glycoproteins. Infect Immun. 1989;57(11):3324–3330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Chong DC, Paton JC, Thorpe CM, Paton AW. Clathrin-dependent trafficking of subtilase cytotoxin, a novel AB5 toxin that targets the endoplasmic reticulum chaperone BiP. Cellular Microbiology. 2008;10(3):795–806. [DOI] [PubMed] [Google Scholar]
- 26.Nagasawa S, Ogura K, Tsutsuki H, et al. Uptake of Shiga-toxigenic Escherichia coli SubAB by HeLa cells requires an actin- and lipid raft -dependent pathway. Cellular microbiology. 2014;16(10):1582–1601. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Sandvig K, van Deurs B. Endocytosis, intracellular transport, and cytotoxic action of Shiga toxin and ricin. Physiological Reviews. 1996;76(4):949–966. [DOI] [PubMed] [Google Scholar]
- 28.Torgersen ML, Skretting G, van Deurs B, Sandvig K. Internalization of cholera toxin by different endocytic mechanisms. Journal of Cell Science. 2001;114(20):3737–3747. [DOI] [PubMed] [Google Scholar]
- 29.Falnes PO, Sandvig K. Penetration of protein toxins into cells. Curr Opin Cell Biol. 2000;12(4):407–413. [DOI] [PubMed] [Google Scholar]
- 30.Lencer WI, Saslowsky D. Raft trafficking of AB5 subunit bacterial toxins. Biochimica et Biophysica Acta (BBA) - Molecular Cell Research. 2005;1746(3):314–321. [DOI] [PubMed] [Google Scholar]
- 31.Plaut RD, Carbonetti NH. Retrograde transport of pertussis toxin in the mammalian cell. Cell Microbiol. 2008;10(5):1130–1139. [DOI] [PubMed] [Google Scholar]
- 32.Sandvig K, Grimmer S, Lauvrak SU, et al. Pathways followed by ricin and Shiga toxin into cells. Histochem Cell Biol. 2002;117(2):131–141. [DOI] [PubMed] [Google Scholar]
- 33.Yu M, Haslam DB. Shiga Toxin Is Transported from the Endoplasmic Reticulum following Interaction with the Luminal Chaperone HEDJ/ERdj3. Infection and Immunity. 2005;73(4):2524–2532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Bellisola G, Fracasso G, Ippoliti R, et al. Reductive activation of ricin and ricin A-chain immunotoxins by protein disulfide isomerase and thioredoxin reductase. Biochem Pharmacol. 2004;67(9):1721–1731. [DOI] [PubMed] [Google Scholar]
- 35.Orlandi PA. Protein-disulfide Isomerase-mediated Reduction of the A Subunit of Cholera Toxin in a Human Intestinal Cell Line. Journal of Biological Chemistry. 1997;272(7):4591–4599. [PubMed] [Google Scholar]
- 36.Spooner RA, Watson PD, Marsden CJ, et al. Protein disulphide-isomerase reduces ricin to its A and B chains in the endoplasmic reticulum. Biochem J. 2004;383(Pt 2):285–293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Tsai B, Rodighiero C, Lencer WI, Rapoport TA. Protein Disulfide Isomerase Acts as a Redox- Dependent Chaperone to Unfold Cholera Toxin. Cell. 2001;104(6):937–948. [DOI] [PubMed] [Google Scholar]
- 38.Schmitz A, Herrgen H, Winkeler A, Herzog V. Cholera Toxin Is Exported from Microsomes by the Sec61p Complex. The Journal of Cell Biology. 2000;148(6):1203–1212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Kahn RA, Gilman AG. Purification of a protein cofactor required for ADP-ribosylation of the stimulatory regulatory component of adenylate cyclase by cholera toxin. J Biol Chem. 1984;259(10):6228–6234. [PubMed] [Google Scholar]
- 40.Bobak DA, Bliziotes MM, Noda M, Tsai SC, Adamik R, Moss J. Mechanism of activation of cholera toxin by ADP-ribosylation factor (ARF): both low- and high-affinity interactions of ARF with guanine nucleotides promote toxin activation. Biochemistry. 1990;29(4):855–861. [DOI] [PubMed] [Google Scholar]
- 41.Le Nours J, Paton AW, Byres E, et al. Structural basis of subtilase cytotoxin SubAB assembly. J Biol Chem. 2013;288(38):27505–27516. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Paton AW, Beddoe T, Thorpe CM, et al. AB5 subtilase cytotoxin inactivates the endoplasmic reticulum chaperone BiP. Nature. 2006;443(7111):548–552. [DOI] [PubMed] [Google Scholar]
- 43.Jobling MG, Yang Z, Kam WR, Lencer WI, Holmes RK. A single native ganglioside GM1-binding site is sufficient for cholera toxin to bind to cells and complete the intoxication pathway. MBio. 2012;3(6). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Lencer WI, Constable C, Moe S, et al. Targeting of cholera toxin and Escherichia coli heat labile toxin in polarized epithelia: role of COOH-terminal KDEL. J Cell Biol. 1995;131(4):951–962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Varki A Essentials of glycobiology. 2nd ed. Cold Spring Harbor, N.Y.: Cold Spring Harbor Laboratory Press; 2009. [PubMed] [Google Scholar]
- 46.Kato K, Ishiwa A. The Role of Carbohydrates in Infection Strategies of Enteric Pathogens. Tropical Medicine and Health. 2015;43(1):41–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Lingwood CA. Glycosphingolipid Functions. Cold Spring Harbor Perspectives in Biology. 2011;3(7):a004788. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Yu RK. Gangliosides: Structure and Analysis In: Ledeen RW, Yu RK, Rapport MM, Suzuki K, eds. Ganglioside Structure, Function, and Biomedical Potential. Boston, MA: Springer US; 1984:39–53. [Google Scholar]
- 49.Chinnapen DJF, Hsieh W-T, te Welscher YM, et al. Lipid-sorting by ceramide structure from plasma membrane to ER for the cholera toxin receptor ganglioside GM1. Developmental cell. 2012;23(3):573–586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Kolter T Ganglioside biochemistry. ISRN Biochem. 2012;2012:506160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Yu RK, Tsai Y-T, Ariga T, Yanagisawa M. Structures, biosynthesis, and functions of gangliosides—An overview. Journal of oleo science. 2011;60(10):537–544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Varki A Loss of N-glycolylneuraminic acid in humans: Mechanisms, consequences, and implications for hominid evolution. American Journal of Physical Anthropology. 2001;116(S33):54–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Schnaar RL, Gerardy-Schahn R, Hildebrandt H. Sialic Acids in the Brain: Gangliosides and Polysialic Acid in Nervous System Development, Stability, Disease, and Regeneration. Physiological Reviews. 2014;94(2):461–518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Daniotti JL, Iglesias-Bartolome R. Metabolic pathways and intracellular trafficking of gangliosides. IUBMB Life. 2011;63(7):513–520. [DOI] [PubMed] [Google Scholar]
- 55.Ikonen E Roles of lipid rafts in membrane transport. Current Opinion in Cell Biology. 2001;13(4):470–477. [DOI] [PubMed] [Google Scholar]
- 56.Simons K, Ikonen E. Functional rafts in cell membranes. Nature. 1997;387(6633):569–572. [DOI] [PubMed] [Google Scholar]
- 57.Vyas AA, Patel HV, Fromholt SE, et al. Gangliosides are functional nerve cell ligands for myelin- associated glycoprotein (MAG), an inhibitor of nerve regeneration. Proc Natl Acad Sci U S A. 2002;99(12):8412–8417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Posse de Chaves E, Sipione S. Sphingolipids and gangliosides of the nervous system in membrane function and dysfunction. FEBS Letters. 2010;584(9):1748–1759. [DOI] [PubMed] [Google Scholar]
- 59.Yoshikawa M, Go S, Takasaki K, et al. Mice lacking ganglioside GM3 synthase exhibit complete hearing loss due to selective degeneration of the organ of Corti. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(23):9483–9488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Yu RK, Tsai Y-T, Ariga T. Functional roles of gangliosides in neurodevelopment--An overview of recent advances. Neurochemical research. 2012;37(6):1230–1244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Kawai H, Allende ML, Wada R, et al. Mice expressing only monosialoganglioside GM3 exhibit lethal audiogenic seizures. J Biol Chem. 2001;276(10):6885–6888. [DOI] [PubMed] [Google Scholar]
- 62.Yamashita T, Wu YP, Sandhoff R, et al. Interruption of ganglioside synthesis produces central nervous system degeneration and altered axon-glial interactions. Proc Natl Acad Sci U S A. 2005;102(8):2725–2730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Rossetto O, Pirazzini M, Montecucco C. Botulinum neurotoxins: genetic, structural and mechanistic insights. Nat Rev Micro. 2014;12(8):535–549. [DOI] [PubMed] [Google Scholar]
- 64.Lacy DB, Tepp W, Cohen AC, DasGupta BR, Stevens RC. Crystal structure of botulinum neurotoxin type A and implications for toxicity. Nat Struct Mol Biol. 1998;5(10):898–902. [DOI] [PubMed] [Google Scholar]
- 65.Blasi J, Chapman ER, Link E, et al. Botulinum neurotoxin A selectively cleaves the synaptic protein SNAP-25. Nature. 1993;365(6442):160–163. [DOI] [PubMed] [Google Scholar]
- 66.Schiavo G, Benfenati F, Poulain B, et al. Tetanus and botulinum-B neurotoxins block neurotransmitter release by proteolytic cleavage of synaptobrevin. Nature. 1992;359(6398):832–835. [DOI] [PubMed] [Google Scholar]
- 67.Neale EA, Bowers LM, Jia M, Bateman KE, Williamson LC. Botulinum Neurotoxin a Blocks Synaptic Vesicle Exocytosis but Not Endocytosis at the Nerve Terminal. The Journal of Cell Biology. 1999;147(6):1249–1260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Sudhof TC, Rizo J. Synaptic vesicle exocytosis. Cold Spring Harbor perspectives in biology. 2011;3(12):10.1101/cshperspect.a005637 a005637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Chen S, Hall C, Barbieri JT. Substrate Recognition of VAMP-2 by Botulinum Neurotoxin B and Tetanus Neurotoxin. The Journal of Biological Chemistry. 2008;283(30):21153–21159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Popoff MR, Poulain B. Bacterial Toxins and the Nervous System: Neurotoxins and Multipotential Toxins Interacting with Neuronal Cells. Toxins. 2010;2(4):683–737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Schengrund C-L, DasGupta BR, Ringler NJ. Binding of Botulinum and Tetanus Neurotoxins to Ganglioside GT1b and Derivatives Thereof. Journal of Neurochemistry. 1991;57(3):1024–1032. [DOI] [PubMed] [Google Scholar]
- 72.Walton KM, Sandberg K, Rogers TB, Schnaar RL. Complex ganglioside expression and tetanus toxin binding by PC12 pheochromocytoma cells. Journal of Biological Chemistry. 1988;263(4):2055–2063. [PubMed] [Google Scholar]
- 73.Rogers TB, Snyder SH. High affinity binding of tetanus toxin to mammalian brain membranes. Journal of Biological Chemistry. 1981;256(5):2402–2407. [PubMed] [Google Scholar]
- 74.Chen C, Baldwin MR, Barbieri JT. Molecular Basis for Tetanus Toxin Coreceptor Interactions. Biochemistry. 2008;47(27):7179–7186. [DOI] [PubMed] [Google Scholar]
- 75.Herreros J, Ng T, Schiavo G. Lipid Rafts Act as Specialized Domains for Tetanus Toxin Binding and Internalization into Neurons. Molecular Biology of the Cell. 2001;12(10):2947–2960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Yeh FL, Dong M, Yao J, et al. SV2 Mediates Entry of Tetanus Neurotoxin into Central Neurons. PLOS Pathogens. 2010;6(11):e1001207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Blum FC, Chen C, Kroken AR, Barbieri JT. Tetanus Toxin and Botulinum Toxin A Utilize Unique Mechanisms To Enter Neurons of the Central Nervous System. Infection and Immunity. 2012;80(5):1662–1669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Lalli G, Schiavo G. Analysis of retrograde transport in motor neurons reveals common endocytic carriers for tetanus toxin and neurotrophin receptor p75NTR. J Cell Biol. 2002;156(2):233–239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Masuyer G, Conrad J, Stenmark P. The structure of the tetanus toxin reveals pH-mediated domain dynamics. EMBO Rep. 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Fischer A, Mushrush DJ, Lacy DB, Montal M. Botulinum neurotoxin devoid of receptor binding domain translocates active protease. PLoS Pathog. 2008;4(12):e1000245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Fischer A, Montal M. Molecular dissection of botulinum neurotoxin reveals interdomain chaperone function. Toxicon. 2013;75:101–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Verderio C, Rossetto O, Grumelli C, Frassoni C, Montecucco C, Matteoli M. Entering neurons: botulinum toxins and synaptic vesicle recycling. EMBO Reports. 2006;7(10):995–999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Kroken AR, Blum FC, Zuverink M, Barbieri JT. Entry of Botulinum neurotoxin subtypes A1 and A2 into neurons. Infection and Immunity. 2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Karalewitz AP, Kroken AR, Fu Z, Baldwin MR, Kim JJ, Barbieri JT. Identification of a unique ganglioside binding loop within botulinum neurotoxins C and D-SA. Biochemistry. 2010;49(37):8117–8126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Spangler BD. Structure and function of cholera toxin and the related Escherichia coli heat-labile enterotoxin. Microbiological Reviews. 1992;56(4):622–647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Sack RB. The discovery of cholera - like enterotoxins produced by Escherichia coli causing secretory diarrhoea in humans. The Indian Journal of Medical Research. 2011;133(2):171–178. [PMC free article] [PubMed] [Google Scholar]
- 87.Backstrom M, Shahabi V, Johansson S, et al. Structural basis for differential receptor binding of cholera and Escherichia coli heat-labile toxins: influence of heterologous amino acid substitutions in the cholera B-subunit. Mol Microbiol. 1997;24(3):489–497. [DOI] [PubMed] [Google Scholar]
- 88.Lai C-Y, Cancedda F, Duffy LK. ADP-ribosyl transferase activity of cholera toxin polypeptide A1 and the effect of limited trypsinolysis. Biochemical and Biophysical Research Communications. 1981;102(3):1021–1027. [DOI] [PubMed] [Google Scholar]
- 89.Northup JK, Sternweis PC, Smigel MD, Schleifer LS, Ross EM, Gilman AG. Purification of the regulatory component of adenylate cyclase. Proceedings of the National Academy of Sciences of the United States of America. 1980;77(11):6516–6520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Cassel D, Pfeuffer T. Mechanism of cholera toxin action: Covalent modification of the guanyl nucleotide-binding protein of the adenylate cyclase system. Proceedings of the National Academy of Sciences of the United States of America. 1978;75(6):2669–2673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Teter K, Allyn RL, Jobling MG, Holmes RK. Transfer of the Cholera Toxin A1 Polypeptide from the Endoplasmic Reticulum to the Cytosol Is a Rapid Process Facilitated by the Endoplasmic Reticulum-Associated Degradation Pathway. Infection and Immunity. 2002;70(11):6166–6171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Kassis S, Hagmann J, Fishman PH, Chang PP, Moss J. Mechanism of action of cholera toxin on intact cells. Generation of A1 peptide and activation of adenylate cyclase. Journal of Biological Chemistry. 1982;257(20):12148–12152. [PubMed] [Google Scholar]
- 93.Bharati K, Ganguly NK. Cholera toxin: A paradigm of a multifunctional protein. The Indian Journal of Medical Research. 2011;133(2):179–187. [PMC free article] [PubMed] [Google Scholar]
- 94.Merritt EA, Sixma TK, Kalk KH, van Zanten BAM, Hol WGJ. Galactose-binding site in Escherichia coli heat-labile enterotoxin (LT) and cholera toxin (CT). Molecular Microbiology. 1994;13(4):745–753. [DOI] [PubMed] [Google Scholar]
- 95.Barra JL, Monferran CG, Balanzino LE, Cumar FA. Escherichia coli heat-labile enterotoxin preferentially interacts with blood group A-active glycolipids from pig intestinal mucosa and A- and B-active glycolipids from human red cells compared to H-active glycolipids. Molecular and Cellular Biochemistry. 1992;115(1):63–70. [DOI] [PubMed] [Google Scholar]
- 96.Horstman AL, Kuehn MJ. Bacterial Surface Association of Heat-labile Enterotoxin through Lipopolysaccharide after Secretion via the General Secretory Pathway. Journal of Biological Chemistry. 2002;277(36):32538–32545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Galván EM, Diema CD, Roth GA, Monferran CG. Ability of Blood Group A-active Glycosphingolipids to Act as Escherichia coli Heat-Labile Enterotoxin Receptors in HT-29 Cells. The Journal of Infectious Diseases. 2004;189(9):1556–1564. [DOI] [PubMed] [Google Scholar]
- 98.Zalem D, Ribeiro JP, Varrot A, Lebens M, Imberty A, Teneberg S. Biochemical and structural characterization of the novel sialic acid-binding site of Escherichia coli heat-labile enterotoxin LT-IIb. Biochemical Journal. 2016;473(21):3923–3936. [DOI] [PubMed] [Google Scholar]
- 99.Berenson CS, Nawar HF, Kruzel RL, Mandell LM, Connell TD. Ganglioside-binding specificities of E. coli enterotoxin LT-IIc: Importance of long-chain fatty acyl ceramide. Glycobiology. 2013;23(1):23–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Melton-Celsa AR. Shiga Toxin (Stx) Classification, Structure, and Function. Microbiology spectrum. 2014;2(2):10.1128/microbiolspec.EHEC-0024-2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Niyogi SK. Shigellosis. Journal of microbiology (Seoul, Korea). 2005;43(2):133–143. [PubMed] [Google Scholar]
- 102.Endo Y, Tsurugi K. RNA N-glycosidase activity of ricin A-chain. Mechanism of action of the toxic lectin ricin on eukaryotic ribosomes. Journal of Biological Chemistry. 1987;262(17):8128–8130. [PubMed] [Google Scholar]
- 103.Tumer NE, Li XP. Interaction of ricin and Shiga toxins with ribosomes. Curr Top Microbiol Immunol. 2012;357:1–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Endo Y, Tsurugi K, Yutsudo T, Takeda Y, Ogasawara T, Igarashi K. Site of action of a Vero toxin (VT2) from Escherichia coli O157:H7 and of Shiga toxin on eukaryotic ribosomes. European Journal of Biochemistry. 1988;171(1–2):45–50. [DOI] [PubMed] [Google Scholar]
- 105.Tesh VL. ACTIVATION OF CELL STRESS RESPONSE PATHWAYS BY SHIGA TOXINS. Cellular microbiology. 2012;14(1):1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Fraser ME, Fujinaga M, Cherney MM, et al. Structure of Shiga Toxin Type 2 (Stx2) from Escherichia coli O157:H7. Journal of Biological Chemistry. 2004;279(26):27511–27517. [DOI] [PubMed] [Google Scholar]
- 107.Cilmi SA, Karalius BJ, Choy W, Smith RN, Butterton JR. Fabry Disease in Mice Protects against Lethal Disease Caused by Shiga Toxin-Expressing Enterohemorrhagic Escherichia coli. The Journal of Infectious Diseases. 2006;194(8):1135–1140. [DOI] [PubMed] [Google Scholar]
- 108.Hanashima T, Miyake M, Yahiro K, et al. Effect of Gb3 in lipid rafts in resistance to Shiga-like toxin of mutant Vero cells. Microbial Pathogenesis. 2008;45(2):124–133. [DOI] [PubMed] [Google Scholar]
- 109.Kovbasnjuk O, Edidin M, Donowitz M. Role of lipid rafts in Shiga toxin 1 interaction with the apical surface of Caco-2 cells. Journal of Cell Science. 2001;114(22):4025–4031. [DOI] [PubMed] [Google Scholar]
- 110.Mukhopadhyay S, Redler B, Linstedt AD. Shiga toxin–binding site for host cell receptor GPP130 reveals unexpected divergence in toxin-trafficking mechanisms. Molecular Biology of the Cell. 2013;24(15):2311–2318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Mayer CL, Leibowitz CS, Kurosawa S, Stearns-Kurosawa DJ. Shiga Toxins and the Pathophysiology of Hemolytic Uremic Syndrome in Humans and Animals. Toxins. 2012;4(11):1261–1287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Sandvig K, Dubinina E, Garred O, et al. Entry of Shiga toxin into cells. Zentralbl Bakteriol. 1993;278(2–3):296–305. [DOI] [PubMed] [Google Scholar]
- 113.Natarajan R, Linstedt AD. A Cycling cis-Golgi Protein Mediates Endosome-to-Golgi Traffic. Molecular Biology of the Cell. 2004;15(11):4798–4806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Mukhopadhyay S, Linstedt AD. Manganese Blocks Intracellular Trafficking of Shiga Toxin and Protects Against Shiga Toxicosis. Science (New York, NY). 2012;335(6066):332–335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Roy CJ, Song K, Sivasubramani SK, Gardner DJ, Pincus SH. Animal Models of Ricin Toxicosis. Current topics in microbiology and immunology. 2012;357:243–257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Tumer NE, Li X-P. Interaction of Ricin and Shiga Toxins with Ribosomes. Current topics in microbiology and immunology. 2012;357:1–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Audi J, Belson M, Patel M, Schier J, Osterloh J. Ricin poisoning: A comprehensive review. JAMA. 2005;294(18):2342–2351. [DOI] [PubMed] [Google Scholar]
- 118.Frankel AE, Burbage C, Fu T, Tagge E, Chandler J, Willingham MC. Ricin Toxin Contains at Least Three Galactose-Binding Sites Located in B Chain Subdomains 1a, ip, and 2y. Biochemistry. 1996;35(47):14749–14756. [DOI] [PubMed] [Google Scholar]
- 119.Montfort W, Villafranca JE, Monzingo AF, et al. The three-dimensional structure of ricin at 2.8 A. J Biol Chem. 1987;262(11):5398–5403. [PubMed] [Google Scholar]
- 120.Magnusson S, Kjeken R, Berg T. Characterization of Two Distinct Pathways of Endocytosis of Ricin by Rat Liver Endothelial Cells. Experimental Cell Research. 1993;205(1):118–125. [DOI] [PubMed] [Google Scholar]
- 121.Simmons BM, Stahl PD, Russell JH. Mannose receptor-mediated uptake of ricin toxin and ricin A chain by macrophages. Multiple intracellular pathways for a chain translocation. Journal of Biological Chemistry. 1986;261(17):7912–7920. [PubMed] [Google Scholar]
- 122.Magnússon S, Berg T. Endocytosis of ricin by rat liver cells in vivo and in vitro is mainly mediated by mannose receptors on sinusoidal endothelial cells. Biochemical Journal. 1993;291(Pt 3):749–755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Alder NN, Shen Y, Brodsky JL, Hendershot LM, Johnson AE. The molecular mechanisms underlying BiP-mediated gating of the Sec61 translocon of the endoplasmic reticulum. The Journal of Cell Biology. 2005;168(3):389–399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Byres E, Paton AW, Paton JC, et al. Incorporation of a non-human glycan mediates human susceptibility to a bacterial toxin. Nature. 2008;456(7222):648–652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Schauer R, Srinivasan GV, Coddeville B, Zanetta J-P, Guerardel Y. Low incidence of N-glycolylneuraminic acid in birds and reptiles and its absence in the platypus. Carbohydrate Research. 2009;344(12):1494–1500. [DOI] [PubMed] [Google Scholar]
- 126.Carbonetti NH. Pertussis toxin and adenylate cyclase toxin: key virulence factors of Bordetella pertussis and cell biology tools. Future microbiology. 2010;5:455–469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Pittman M The concept of pertussis as a toxin-mediated disease. Pediatr Infect Dis. 1984;3(5):467–486. [DOI] [PubMed] [Google Scholar]
- 128.Kaslow HR, Burns DL. Pertussis toxin and target eukaryotic cells: binding, entry, and activation. The FASEB Journal. 1992;6(9):2684–2690. [DOI] [PubMed] [Google Scholar]
- 129.Sekura RD, Fish F, Manclark CR, Meade B, Zhang YL. Pertussis toxin. Affinity purification of a new ADP-ribosyltransferase. Journal of Biological Chemistry. 1983;258(23):14647–14651. [PubMed] [Google Scholar]
- 130.Stein PE, Boodhoo A, Armstrong GD, Cockle SA, Klein MH, Read RJ. The crystal structure of pertussis toxin. Structure. 1994;2(1):45–57. [DOI] [PubMed] [Google Scholar]
- 131.Stein PE, Boodhoo A, Armstrong GD, et al. Structure of a pertussis toxin-sugar complex as a model for receptor binding. Nat Struct Mol Biol. 1994;1(9):591–596. [DOI] [PubMed] [Google Scholar]
- 132.Finck-Barbancon V, Barbieri JT. Preferential processing of the S1 subunit of pertussis toxin that is bound to eukaryotic cells. Mol Microbiol. 1996;22(1):87–95. [DOI] [PubMed] [Google Scholar]
- 133.Nencioni L, Pizza MG, Volpini G, De Magistris MT, Giovannoni F, Rappuoli R. Properties of the B oligomer of pertussis toxin. Infection and Immunity. 1991;59(12):4732–4734. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Millen SH, Lewallen DM, Herr AB, Iyer SS, Weiss AA. Identification and Characterization of the Carbohydrate Ligands Recognized by Pertussis Toxin through Glycan Microarray and Surface Plasmon Resonance. Biochemistry. 2010;49(28):5954–5967. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Hausman SZ, Burns DL. Binding of pertussis toxin to lipid vesicles containing glycolipids. Infect Immun. 1993;61(1):335–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Brennan MJ, David JL, Kenimer JG, Manclark CR. Lectin-like binding of pertussis toxin to a 165-kilodalton Chinese hamster ovary cell glycoprotein. Journal of Biological Chemistry. 1988;263(10):4895–4899. [PubMed] [Google Scholar]
- 137.Naylor CE, Eaton JT, Howells A, et al. Structure of the key toxin in gas gangrene. Nat Struct Biol. 1998;5(8):738–746. [DOI] [PubMed] [Google Scholar]
- 138.Sakurai J, Nagahama M, Oda M. Clostridium perfringens alpha-toxin: characterization and mode of action. J Biochem. 2004;136(5):569–574. [DOI] [PubMed] [Google Scholar]
- 139.Caplan ES, Kluge RM. Gas gangrene: Review of 34 cases. Archives of Internal Medicine. 1976;136(7):788–791. [DOI] [PubMed] [Google Scholar]
- 140.Oda M, Terao Y, Sakurai J, Nagahama M. Membrane-Binding Mechanism of Clostridium perfringens Alpha-Toxin. Toxins. 2015;7(12):5268–5275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Oda M, Kabura M, Takagishi T, et al. Clostridium perfringens alpha-toxin recognizes the GM1a-TrkA complex. J Biol Chem. 2012;287(39):33070–33079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Takagishi T, Oda M, Kabura M, et al. Clostridium perfringens Alpha-Toxin Induces Gm1a Clustering and Trka Phosphorylation in the Host Cell Membrane. PLOS ONE. 2015;10(4):e0120497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.den Akker Fv, Sarfaty S, Twiddy EM, Connell TD, Holmes RK, Hol WGJ. Crystal structure of a new heat-labile enterotoxin, LT-IIb. Structure. 1996;4(6):665–678. [DOI] [PubMed] [Google Scholar]
- 144.Merritt EA, Sarfaty S, van den Akker F, L’Hoir C, Martial JA, Hol WG. Crystal structure of cholera toxin B-pentamer bound to receptor GM1 pentasaccharide. Protein Sci. 1994;3(2):166–175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Murzin AG. OB(oligonucleotide/oligosaccharide binding)-fold: common structural and functional solution for non-homologous sequences. The EMBO Journal. 1993;12(3):861–867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Merritt EA, Sarfaty S, van den Akker F, L’Hoir C, Martial JA, Hol WG. Crystal structure of cholera toxin B-pentamer bound to receptor GM1 pentasaccharide. Protein Science: A Publication of the Protein Society. 1994;3(2):166–175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Sixma TK, Kalk KH, van Zanten BAM, et al. Refined Structure of Escherichia coli Heat-labile Enterotoxin, a Close Relative of Cholera Toxin. Journal of Molecular Biology. 1993;230(3):890–918. [DOI] [PubMed] [Google Scholar]
- 148.Mudrak B, Kuehn MJ. Heat-labile enterotoxin: beyond G(m1) binding. Toxins (Basel). 2010;2(6):1445–1470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Loosmore S, Zealey G, Cockle S, et al. Characterization of pertussis toxin analogs containing mutations in B-oligomer subunits. Infection and Immunity. 1993;61(6):2316–2324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Ling H, Boodhoo A, Hazes B, et al. Structure of the Shiga-like Toxin I B-Pentamer Complexed with an Analogue of Its Receptor Gb3. Biochemistry. 1998;37(7):1777–1788. [DOI] [PubMed] [Google Scholar]
- 151.Yao J, Nellas RB, Glover MM, Shen T. Stability and Sugar Recognition Ability of Ricin-like Carbohydrate Binding Domains. Biochemistry. 2011;50(19):4097–4104. [DOI] [PubMed] [Google Scholar]
- 152.Sphyris N, Lord JM, Wales R, Roberts LM. Mutational analysis of the Ricinus lectin B-chains. Galactose-binding ability of the 2 gamma subdomain of Ricinus communis agglutinin B-chain. J Biol Chem. 1995;270(35):20292–20297. [DOI] [PubMed] [Google Scholar]
- 153.Wales R, Richardson PT, Roberts LM, Lord JM. Recombinant ricin B chain fragments containing a single galactose binding site retain lectin activity. Arch Biochem Biophys. 1992;294(1):291–296. [DOI] [PubMed] [Google Scholar]
- 154.Emsley P, Fotinou C, Black I, et al. The Structures of the HC Fragment of Tetanus Toxin with Carbohydrate Subunit Complexes Provide Insight into Ganglioside Binding. Journal of Biological Chemistry. 2000;275(12):8889–8894. [DOI] [PubMed] [Google Scholar]
- 155.Stenmark P, Dupuy J, Imamura A, Kiso M, Stevens RC. Crystal Structure of Botulinum Neurotoxin Type A in Complex with the Cell Surface Co-Receptor GT1b—Insight into the Toxin–Neuron Interaction. PLOS Pathogens. 2008;4(8):e1000129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Uppalapati SR, Kingston JJ, Qureshi IA, Murali HS, Batra HV. In Silico, In Vitro and In Vivo Analysis of Binding Affinity between N and C-Domains of Clostridium perfringens Alpha Toxin. PLOS ONE. 2013;8(12):e82024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Burke P, Schooler K, Wiley HS. Regulation of Epidermal Growth Factor Receptor Signaling by Endocytosis and Intracellular Trafficking. Molecular Biology of the Cell. 2001;12(6):1897–1910. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Wang X, Rahman Z, Sun P, et al. Ganglioside Modulates Ligand Binding to the Epidermal Growth Factor Receptor. Journal of Investigative Dermatology. 2001;116(1):69–76. [DOI] [PubMed] [Google Scholar]
- 159.Toledo MS, Suzuki E, Handa K, Hakomori S. Effect of Ganglioside and Tetraspanins in Microdomains on Interaction of Integrins with Fibroblast Growth Factor Receptor. Journal of Biological Chemistry. 2005;280(16):16227–16234. [DOI] [PubMed] [Google Scholar]
- 160.Yoon S-J, Nakayama K-i, Hikita T, Handa K, Hakomori S-i. Epidermal growth factor receptor tyrosine kinase is modulated by GM3 interaction with N-linked GlcNAc termini of the receptor. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(50):18987–18991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Duchemin A-M, Ren Q, Mo L, Neff NH, Hadjiconstantinou M. GM1 ganglioside induces phosphorylation and activation of Trk and Erk in brain. Journal of Neurochemistry. 2002;81(4):696–707. [DOI] [PubMed] [Google Scholar]
- 162.Gil C, Chaïb-Oukadour I, Pelliccioni P, Aguilera J. Activation of signal transduction pathways involving trkA, PLCγ−1, PKC isoforms and ERK-1/2 by tetanus toxin. FEBS Letters. 2000;481(2):177–182. [DOI] [PubMed] [Google Scholar]
- 163.Gil C, Chaib-Oukadour I, Aguilera J. C-terminal fragment of tetanus toxin heavy chain activates Akt and MEK/ERK signalling pathways in a Trk receptor-dependent manner in cultured cortical neurons. Biochem J. 2003;373(Pt 2):613–620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Gil C, Chaib-Oukadour I, Blasi J, Aguilera J. HC fragment (C-terminal portion of the heavy chain) of tetanus toxin activates protein kinase C isoforms and phosphoproteins involved in signal transduction. Biochem J. 2001;356(Pt 1):97–103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Yamamoto N, Matsubara T, Sato T, Yanagisawa K. Age-dependent high-density clustering of GM1 ganglioside at presynaptic neuritic terminals promotes amyloid beta-protein fibrillogenesis. Biochim Biophys Acta. 2008;1778(12):2717–2726. [DOI] [PubMed] [Google Scholar]
- 166.Mutoh T, Tokuda A, Guroff G, Fujiki N. The Effect of the B Subunit of Cholera Toxin on the Action of Nerve Growth Factor on PC12 Cells. Journal of Neurochemistry. 1993;60(4):1540–1547. [DOI] [PubMed] [Google Scholar]
- 167.Lauvrak SU, Walchli S, Iversen TG, et al. Shiga toxin regulates its entry in a Syk-dependent manner. Mol Biol Cell. 2006;17(3):1096–1109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Utskarpen A, Massol R, van Deurs B, Lauvrak SU, Kirchhausen T, Sandvig K. Shiga Toxin Increases Formation of Clathrin-Coated Pits through Syk Kinase. PLOS ONE. 2010;5(7):e10944. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Sonnino S, Mauri L, Chigorno V, Prinetti A. Gangliosides as components of lipid membrane domains. Glycobiology. 2007;17(1):1R–13R. [DOI] [PubMed] [Google Scholar]
- 170.Arab S, Lingwood CA. Influence of phospholipid chain length on verotoxin/globotriaosyl ceramide binding in model membranes: comparison of a supported bilayer film and liposomes. Glycoconj J. 1996;13(2):159–166. [DOI] [PubMed] [Google Scholar]
- 171.Falguieres T, Romer W, Amessou M, et al. Functionally different pools of Shiga toxin receptor, globotriaosyl ceramide, in HeLa cells. FEBS J. 2006;273(22):5205–5218. [DOI] [PubMed] [Google Scholar]
- 172.Gallegos KM, Conrady DG, Karve SS, Gunasekera TS, Herr AB, Weiss AA. Shiga Toxin Binding to Glycolipids and Glycans. PLOS ONE. 2012;7(2):e30368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Smith DC, Sillence DJ, Falguières T, et al. The Association of Shiga-like Toxin with Detergent-resistant Membranes Is Modulated by Glucosylceramide and Is an Essential Requirement in the Endoplasmic Reticulum for a Cytotoxic Effect. Molecular Biology of the Cell. 2006;17(3):1375–1387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Wolf AA, Jobling MG, Wimer-Mackin S, et al. Ganglioside Structure Dictates Signal Transduction by Cholera Toxin and Association with Caveolae-like Membrane Domains in Polarized Epithelia. The Journal of Cell Biology. 1998;141(4):917–927. [DOI] [PMC free article] [PubMed] [Google Scholar]




