Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2012 Feb 6;590(Pt 6):1357–1368. doi: 10.1113/jphysiol.2011.225276

Copper-dependent regulation of NMDA receptors by cellular prion protein: implications for neurodegenerative disorders

Peter K Stys 1, Haitao You 2, Gerald W Zamponi 2
PMCID: PMC3382327  PMID: 22310309

Abstract

Abstract

N-Methyl-d-aspartate (NMDA) receptors mediate a wide range of important nervous system functions. Conversely, excessive NMDA receptor activity leads to cytotoxic calcium overload and neuronal damage in a wide variety of CNS disorders. It is well established that NMDA receptors are tightly regulated by a number of cell signalling pathways. Recently, it has been shown that NMDA receptor activity is modulated by cellular prion protein (PrPC) in a copper-dependent manner. Here we give an overview of the current state of knowledge concerning the novel concept of potent modulation of this receptor's kinetics by copper ions, and the interplay between NMDA receptors and PrPC in the context of neurological diseases such as Alzheimer's disease, epilepsy, pain and depression.


Drs Gerald Zamponi and Peter Stys work at the University of Calgary's Hotchkiss Brain Institute. Dr Zamponi is the Head of the Department of Physiology and Pharmacology and a basic neuroscientist. His interest lies in the molecular physiology of voltage and ligand gated ion channels with an emphasis on pain and neurodegenerative disorders. Dr Stys is a neurologist and clinician scientist with expertise in multiphoton imaging techniques and a research interest in multiple sclerosis and neurodegeneration. Through a collaborative effort Drs Zamponi and Stys have discovered a new mechanism by which Aβ mediates neuronal degeneration via a copper dependent dysregulation of NMDA receptors.

graphic file with name tjp0590-1357-m1.jpg

NMDA receptor function

N-Methyl-d-aspartate (NMDA) receptors are one of the major classes of ionotropic glutamate receptors in the mammalian brain (McBain & Mayer, 1994). Their activation by glutamate or the synthetic agonist NMDA, together with the co-agonist glycine (or d-serine) (Shleper et al. 2005), results in the opening of a non-selective cation channel that mediates entry of calcium and sodium ions into the cytosol. At the synapse, glutamate release from presynaptic terminals diffuses across the synaptic cleft to activate these receptors, which in turn depolarize the postsynaptic membrane to induce an excitatory postsynaptic potential (EPSP). In addition, there are extrasynaptic NMDA receptors that are activated by glutamate spillover (Asztely et al. 1997; Lozovaya et al. 1999; Diamond, 2001). It is well established that NMDA receptors play a critical function in learning and memory (Herron et al. 1986; Collingridge, 1987; MacDonald et al. 2006). However, altered expression or function of these receptors is likely to play a role in the pathophysiology of a wide variety of CNS disorders including ischaemia, epilepsy, many neurodegenerative diseases and even neuropsychiatric disorders such as schizophrenia (Lipton & Rosenberg, 1994; Aarts et al. 2002; Loftis & Janowsky, 2003). The fact that NMDA receptors fulfil key physiological functions poses a dilemma with regard to designing therapeutics targeted towards these receptors, although a balance can be struck as evidenced by the Alzheimer's drug memantine, which is an NMDA receptor antagonist with relatively modest side effects (Lipton, 2006; Olivares et al. 2011). At negative voltages near the resting membrane potential, most NMDA receptors are tonically inhibited by extracellular magnesium ions (Nowak et al. 1984). Membrane depolarization dislodges magnesium from its binding site in the NMDA receptor pore, thus conferring a voltage dependence on receptor activity (Mayer et al. 1984). In addition to magnesium, zinc ions also bind to the receptors outside of the pore to modulate activity (Peters et al. 1987). As we will outline below, another metal ion that is emerging as an important NMDA receptor modulator is copper.

NMDA receptors activate (i.e. open) in response to ligand binding, and they deactivate (i.e. close) upon ligand unbinding (Vance et al. 2011). During prolonged application of agonist, these receptors undergo desensitization (Mayer et al. 1989), a process conceptually similar to voltage-dependent inactivation in voltage-gated calcium and sodium channels (Stotz et al. 2004), and designed to limit toxic calcium overload of cells during periods of prolonged glutamate elevations. Receptor desensitization is modulated by the aforementioned co-agonist glycine, such that increasing concentrations of this ligand produce a dramatic slowing of receptor desensitization kinetics (Mayer et al. 1989; Vyklicky et al. 1990). The physiological importance of NMDA receptor desensitization is underscored by the observation that application of high glycine concentrations is toxic to neurons, and blocking glycine reuptake increases NMDA receptor mediated neuronal excitability (Chen et al. 2003). Hence, receptor kinetics must be tightly controlled to maintain critical NMDA receptor functionality while preventing receptor hyperactivity and cellular damage.

NMDA receptor structure and subunit composition

NMDA receptors are heterotetramers that contain two obligatory GluN1 subunits (formerly known as NR1; Collingridge et al. 2009) and various combinations of GluN2A–D (formerly NR2A–D) and GluN3A–B (formerly NR3A–B) subunits (Paoletti & Neyton, 2007). Each of these subunits share a common transmembrane topology with a large extracellular N-terminal domain, and four membrane helices, as well as a large cytoplasmic C-terminal region (see Fig. 1) (Paoletti, 2011). In mammals, there is only one gene encoding GluN1, although different isoforms can be generated through alternative splicing (Tingley et al. 1993). The GluN1 subunit contains the binding site for the co-agonist glycine and d-serine which interact at a site formed by the N-terminal domain and the extracellular linker connecting the third and fourth transmembrane domains (for review see Paoletti, 2011). There are four different types of GluN2 subunits (GluN2A, B, C and D) which can be incorporated with the two obligatory GluN1 subunits into the assembled NMDA receptor complex, or with GluN3 to form triheteromers (Low & Wee, 2010; Rauner & Kohr, 2011). The GluN2 subunit contains the glutamate binding site, which, similarly to the GluN1 glycine binding site, is formed by interactions between the N-terminal domain and the loop between M3 and M4. The GluN2 subunit also contains the zinc interaction site which is located within the N-terminal domain (Choi & Lipton, 1999). The GluN3 subunits (i.e. GluN3A and GluN3B) do not bind glutamate, but instead interact with glycine, thus giving rise to a ‘glycine-only NMDA receptor’ when assembled with GluN1 in the absence of GluN2 (Chatterton et al. 2002; Piña-Crespo et al. 2010). In the context of this article, these glutamate-independent receptors are not considered further. In a fully assembled receptor, the pore region is formed by residues within the M2 regions which function similarly to the p-loops found in other types of ion channels (Kupper et al. 1996). Specific amino acid residues in the pore region ensure selectivity for cations, and the ability to bind magnesium ions or blockers such as MK801 (Mori et al. 1992).

Figure 1. NMDA receptor structure and kinetics.

Figure 1

A, transmembrane topology of a basic NMDA receptor subunit (i.e. GluN1, GluN2 or GluN3). Note the large N-terminus region that is involved in ligand binding. B, schematic representation of an NMDA receptor. One GluN1 and one GluN2 subunit are shown, although the fully assembled receptor complex contains two GluN1 and two GluN2 (or GluN3) subunits. C, effect of glycine on NMDA receptor desensitization. Two current recordings obtained from a wild-type mouse hippocampal neuron are depicted and scaled to overlap at peak. Both traces were evoked by 500 μm NMDA, the traces shown in black and grey were obtained, respectively, in the presence of 300 nm and 10 μm glycine.

The different types of GluN2 subunits generate receptor complexes with distinct functional and pharmacological properties (Loftis & Janowsky, 2003; Cull-Candy & Leszkiewicz, 2004; Paoletti & Neyton, 2007; Paoletti, 2011). As demonstrated in heterologous expression systems, GluN2 subunits govern a number of important NMDA receptor properties, including deactivation and desensitization rates, glycine and glutamate affinity, maximal open probability and susceptibility to magnesium block (reviewed in Paoletti, 2011). Furthermore, in neurons, the subcellular distribution of the receptors is GluN2 subtype dependent with GluN2A-containing receptors being localized to synaptic sites, whereas GluN2B containing receptors are found extrasynaptically (Liu et al. 2007), although this arrangement is not absolute (Thomas et al. 2006). At least in rodents, the expression of GluN2 subunits is thought to change with age, thus adding further complexity and suggesting a specific role of GluN2 subunits at various stages of neuronal development (Loftis & Janowsky, 2003).

NMDA receptors do not operate in isolation. They are tightly regulated by second messengers such as src and fyn kinases and CamKII which act on the large cytoplasmic C-terminal domain of the receptor (Xu et al. 2008; Sanhueza et al. 2011; Trepanier et al. 2012). Furthermore, NMDA receptors interact with certain types of G-protein coupled receptors (such as dopamine receptors) to form macromolecular signalling complexes (Lee et al. 2002). Recent work from our laboratory has revealed that NMDA receptors interact with, and are regulated by, cellular prion protein (PrPC) (Khosravani et al. 2008; You et al. 2012). Below, we will focus on the effect of PrPC on NMDA receptor function, and its possible pathophysiological roles.

Structure and physiological function of PrPC

Before normal cleavage, PrPC is a 254 amino acid protein that is ubiquitously expressed and has highly conserved homologues even in simple organisms such as yeast. It contains a 22 amino acid N-terminal signal peptide, followed, depending on the species, by four to five octapeptide repeats contained within an unstructured N-terminal region (Riesner, 2003; Aguzzi & Heikenwalder, 2006). The C-terminal half of the protein contains three α-helices and a glycophosphatidylinositol (GPI) anchor at position 231 that mediates an association with the extracellular leaflet of the plasma membrane (Aguzzi & Heikenwalder, 2006). The octapeptide repeats contain up to five copper binding sites with affinities varying from the femtomolar to the micromolar range (Jackson et al. 2001; Burns et al. 2003). Conversion of PrPC into the pathological β-sheet-rich scrapie conformation (i.e. PrPSc) has been associated with neurological disorders such as Creutzfeldt–Jakob disease (Kingsbury et al. 1983; Palmer et al. 1991; Mallucci et al. 1999), bovine spongiform encephalopathy (BSE) (Hill et al. 1997; Prusiner, 1997), and Kuru (Gajdusek, 1977). Once misfolded, PrPSc acts as a chaperone that triggers the misfolding of normal PrPC, thus leading to a progressive accumulation of PrPSc (Aguzzi et al. 2008). This in turn gives rise to aggregates that may cause massive neuronal degeneration that ultimately culminate in death of the subject (Caughey et al. 2009). Therefore, much effort has focused on understanding the role and function of PrPSc, and the molecular and biochemical mechanisms that trigger misfolding of normal PrPC. In contrast, the normal physiological function of PrPC remains incompletely understood (Linden et al. 2008). Interestingly, copper, likely acting via the copper-binding octarepeats, induces significant conformational changes to PrPC (Wong et al. 2003), which may have important implications for the regulation of NMDARs (see below). There is also accumulating evidence that PrPC may not only play an important role in neuroprotection (Khosravani et al. 2008), but also act as a conduit for compromised neuronal function and ultimately degeneration in disorders not typically considered prionopathies, such as Alzheimer's disease (Lauren et al. 2009).

Mice lacking PrPC display a surprisingly mild behavioural phenotype (Bueler et al. 1992; for review see Steele et al. 2007) unless subjected to an insult. Under normal circumstances, these mice reportedly show slight deficits in spatial learning (Nishida et al. 1997; Valenti et al. 2001; Criado et al. 2005) and at an advanced age, peripheral nerve demyelination (Bremer et al. 2010). Reports based on brain slice recordings suggest altered long term potentiation (Collinge et al. 1994; Johnston et al. 1998; Curtis et al. 2003; Rangel et al. 2009), consistent with these memory deficits. Notably, PrPC null mice are resistant to infectious prions, because these mice lack the template for further conversion of PrPC into PrPSc and accumulation of the latter (Sailer et al. 1994). Along these lines, knockdown of PrPC prevents PrPSc toxicity in rodents (White et al. 2008; White & Mallucci, 2009). However, there is increasing evidence that the absence of PrPC can in fact be harmful. For example, PrPC null mice show greater mortality following chemically nduced (i.e. pentylenetetrazole and pilocarpine) seizures, and a greater incidence of seizures in the kainic acid model (Walz et al. 1999), although these findings have been disputed in a more recent study (Ratte et al. 2011). PrPC null mice display enhanced and prolonged neuroinflammation in a model of experimental autoimmune encephalomyelitis (EAE), whereas overexpression of PrPC was protective in the EAE model (Tsutsui et al. 2008). Moreover, silencing of PrPC has been shown to accelerate CNS autoimmune disease in a T-cell-dependent manner (Hu et al. 2010). Finally, recent findings from our laboratory have revealed that PrPC null mice display depressive-like behaviour (Gadotti et al. 2012), and exhibit heightened nociception and increased inflammatory pain, as well as central pain sensitization (Gadotti & Zamponi, 2011). Interestingly, both of these behaviours could be abrogated by treatment with the NMDA receptor antagonist MK-801, suggesting the possibility that NMDA receptor activity may be enhanced by the absence of PrPC. Together, these findings indicate that PrPC may have multiple neuroprotective roles, thus perhaps explaining why this protein is so widely expressed.

PrPC regulation of NMDA receptors

PrPC appears to regulate the function of a wide range of different ion channels and receptors. For example, L-type calcium channel amplitude is decreased when the channels are exposed to the copper binding domain of recombinant PrPC (Korte et al. 2003). Knockout of PrPC results in a reduction of calcium activated potassium currents in cerebellar Purkinje cells (Herms et al. 2001), although it is not clear if this is a direct action on the potassium channel, or an indirect effect mediated via altered calcium entry. Along these lines, application of a 20 amino acid central fragment of PrPC (residues 106–126) triggers the activation of potassium currents in basal forebrain neurons (Alier et al. 2010). A subsequent report revealed that PrPC interacts with the potassium channel tetramerization domain (Huang et al. 2012), together indicating that PrPC is an important regulator of various types of potassium channels. A recent study revealed that PrPC interacts with mGLuR1 and mGluR5 receptors to stimulate ERK1/2 activation (Beraldo et al. 2011) suggesting that PrPC can regulate G protein-coupled receptor-mediated cell signalling. Ligand-gated channels are also regulated by PrPC. For example, the α7-nicotinic receptor activity is regulated by stress-inducible-protein-1 via PrPC (Beraldo et al. 2010), and PrPC is known to regulate kainate receptor expression (Rangel et al. 2007). Altogether these examples illustrate that PrPC is a promiscuous protein that has the propensity to regulate signalling by plasma membrane receptors and channels (for a more detailed review, see Linden et al. 2008).

It may thus not be surprising to note that NMDA receptor activity is subject to modulation by PrPC and PrPSc. Indeed, in addition to the in vivo data mentioned in the previous section, regulation of NMDA receptor function by PrPC is also supported by findings that MK801 protects from increased cell death in PrP null mice subjected to kainate excitotoxicity (Rangel et al. 2007), presumably because the kainate-induced membrane depolarization leads to a secondary activation of NMDA receptors. Furthermore, neuronal cultures infected with PrPSc show increased survival in the presence of MK801 (Muller et al. 1993).

A recent study from our group examined the cellular basis for these effects (Khosravani et al. 2008). In field potential recordings from hippocampal slices from PrP null mice there was an increase in the number of population spikes compared with those seen in WT slices. This effect was exacerbated upon removal of magnesium ions, and normalized in the presence of the NMDA receptor blocker amino-5-phosphonovaleric acid. This hyperexcitability seen with PrP null slices fits with the increased susceptibility of PrPC null mice to seizure-like behaviour. Miniature synaptic NMDA currents were also enhanced both in amplitude and duration, and whole cell NMDA currents in cultured pyramidal neurons showed greater current amplitude and dramatically slowed deactivation kinetics. Consistent with NMDA receptor hyperfunction, direct injection of NMDA into PrP null mouse brains resulted in dramatically increased lesion size compared to WT animals (Khosravani et al. 2008). The slowed deactivation kinetics appear reminiscent of what is observed with transiently expressed NMDA receptors containing the GluN2D subunit (Cull-Candy & Leszkiewicz, 2004), perhaps suggesting that the absence of PrPC may lead to altered NMDA receptor subunit composition. However, although GluN2D siRNA treatment resulted in smaller, faster deactivating currents, it is unlikely that PrP null mouse neurons contain pure GluN1/GluN2D receptors, because whole cell currents could still be blocked by a combination of GluN2A and GluN2B blockers (H. You and G. W. Zamponi, unpublished observations). Rather, we suspect that in the absence of PrPC, there is an increased proportion of heteromeric GluN1/GluN2(A or B)/GluN2D receptors which may give rise to altered deactivation kinetics, thus potentially reconciling our findings. The observation that synaptic NMDA currents are upregulated may explain the increased pain phenotype in PrP null mice, as NMDA receptors expressed in the dorsal horn of the spinal cord are similarly upregulated during central pain sensitization (i.e. ‘wind-up’) (Woolf & Thompson, 1991; Salter, 2005; Liu et al. 2008).

As noted above, NMDA receptor inhibitors prevent cell death of cultured neurons exposed to PrPSc (Muller et al. 1993). In principle, this could be explained by one of three different scenarios. First, exposure to PrPSc may cause the misfolding of endogenous PrPC, thus leading to a de facto PrPC knockdown and increased NMDA receptor activity. Indeed, a reduction in PrPC levels has been reported in mouse models of prion disease (Griffin et al. 2007). Alternatively, it is conceivable that PrPSc could directly stimulate NMDA receptor function. Finally, it is possible that PrPSc may be unable to inhibit NMDA receptor activity, thereby acting as a dominant negative inhibitor of PrPC action on the receptor complex.

Copper modulates NMDA receptors via PrPC

It is well established that PrPC is a high affinity copper binding protein (Brown et al. 1997; Jackson et al. 2001; Brown & Sassoon, 2002); however the physiological role of the copper binding sites on PrPC has been elusive. We have recently shown that chelation of copper ions potently regulates native NMDA receptors in rat and mouse hippocampal neurons (You et al. 2012). As noted earlier, glycine acts as a potent co-agonist at the GluN1 receptor and a negative regulator of NMDA receptor desensitization. Many investigators in the NMDA receptor field conduct recordings in high concentrations of glycine as an experimental manipulation to boost current responses. Interestingly, both glycine and d-serine chelate copper ions (Martin et al. 1971), therefore the balance between applied agonist concentration and ambient copper levels becomes critically important. For this reason, we conducted our experiments using a physiological glycine concentration that is thought to reflect the submicromolar levels normally present at the synapse (Supplisson & Roux, 2002; Yang & Svensson, 2008). When neurons were exposed to 300 nm glycine/500 μm NMDA, currents in cultured pyramidal cells desensitized almost completely (Fig. 1C, black trace). Exposure of neurons to selective copper chelators such as bathocuproine sulfonate (BCS) or cuprizone induced a persistent non-desensitizing current component that could be overcome by addition of excess copper ions. Interestingly, in 300 nm glycine (without copper chelation), neurons from PrP null mice behaved exactly like BCS treated WT neurons; moreover the effects of PrPC knockout and BCS were not additive, suggesting that the effects of copper on NMDA receptor desensitization were mediated via PrPC. Furthermore, PrPC could be immunoprecipitated with GluN1, and the strength of this interaction was greatly weakened upon copper chelation. This, together with the observation that the glycine affinity for the receptor was enhanced in PrPC null mice (or upon application of BCS) led us to propose a model in which copper-dependent allosteric interactions between PrPC and the GluN1 subunit regulate the affinity of the receptor complex for the co-agonist glycine, in this manner leading to non-desensitizing currents at any given glycine concentration (Fig. 2). This then suggests that perhaps a key physiological role of the copper binding sites on PrPC is to regulate NMDA receptors. However, even in the absence of PrPC, micromolar concentrations of copper still speed NMDA receptor desensitization (You et al. 2012; see also Vlachova et al. 1996), indicating the presence of a second copper modulation mechanism that operates independently of PrPC. One possibility may be an as-yet-unidentified copper binding site on the receptor, as was discovered for zinc (Rachline et al. 2005). The intrinsic mechanisms that give rise to persistent currents in NMDA receptors upon copper chelation/deletion of PrPC are not known. One attractive possibility is a switch in modal gating of the receptor, as described recently in a study by Zhang et al. (2008). It is also interesting to note that GluN2D-containing receptors show little desensitization and display the highest glycine affinity of all NMDA receptor subtypes (Erreger et al. 2007; Chen et al. 2008). However, given that the effects of copper chelation occurred rapidly, a switch in NMDA receptor subunit composition is unlikely to account for the persistent currents.

Figure 2. Model depicting the effects of copper and PrPC on NMDA receptor function.

Figure 2

A, under normal circumstances, PrPC in its copper-loaded state exists in a complex with the NMDA receptor, reducing the affinity of the co-agonist glycine for the receptor. At physiological levels of glycine (1 μm or lower), this results in strong desensitization of the receptor (see current trace obtained from a cultured hippocampal neuron). B, knockout of the gene encoding PrPC results in a receptor with higher affinity for glycine, thus leading to persistent non-desensitizing currents and overall greater accumulation of calcium in the cell. C, chelation of copper ions by the selective chelator bathocuproine sulfonate or by Aβmonomers causes a weakening of the association between PrPC and the NMDA receptor, thus producing currents similar to those seen in PrPC null mouse neurons. D, Aβoligomers bind directly to PrPC to disrupt PrPC regulation of NMDA receptor activity, producing the same phenotype as that shown in panels B and C.

The regulation of NMDA receptors by copper ions deserves additional discussion. First, BCS might be a selective chelator of Cu+ over Cu2+ whereas PrPC is thought to specifically bind Cu2+. However, the interactions between BCS and the two oxidation states of copper are non-trivial, with BCS able to bind Cu2+ with high affinity to form bis µCu(BCS)2½2− complexes (Al-Shatti et al. 1981; Sayre, 1996) or the protonated form of BCS µH·BCS½ associates with Cu2+ to form complex species such as µCu2+ (BCS) (H2O)x½ and a proton (Xiao et al. 2011); in a pH buffer, this proton will be buffered driving the reaction in favour of complete chelation of all Cu2+ by BCS (or more precisely, by its mono-protonated form). The fact that the canonical Cu2+-selective chelator cuprizone (Peterson & Bollier, 1955) yielded identical results supports the idea that BCS indeed acted as a high-affinity copper chelator independent of this metal's oxidation state.

There is a large body of literature on the copper concentrations in CSF, with published values ranging widely from 0.12 to 7 μm (Kanabrocki et al. 1964; Agarwal & Henkin, 1982). In the whole brain, levels are around 5 μg (g wet weight)-1 (∼80 μm equivalent) (Warren et al. 1960). The problem with determining CSF or whole brain copper levels is that almost all of this metal is bound to proteins and amino acids; thus the copper available for binding to PrPC is determined not so much by the measured total concentrations, but rather by the relative affinities of PrPC vis-à-vis those of other copper binding species. Measured copper in our culture media was around 100 nm although other media range up to the low micromolar level. The resting copper concentration in the synaptic cleft where the NMDA receptors are located is also estimated to be ∼1 μm, though transient peak levels may approach 250 μm (reviewed in Millhauser, 2007). Given the ability of copper to potently modulate NMDA receptor kinetics and limit agonist-induced inward current and therefore Ca2+ loads (see above), it is plausible that ambient copper in culture media and brain extracellular space functions to limit NMDA receptor-dependent excitotoxicity. In line with this prediction, treatment of cultured neurons with copper chelators causes cell death that can be prevented by the NMDA receptor antagonist 5,7-dichlorokynurenic acid or by supplying excess copper (You et al. 2012), underscoring the potential key importance of trace copper in cell culture media and experimental perfusing solutions. Neurons have active copper transport mechanisms such as the copper ATPases (Niciu et al. 2006; Veldhuis et al. 2009) that, along with copper binding proteins such as PrPC, ensure proper copper homeostasis in the brain. When cultured neurons or brain slices are placed in recording chambers filled with external solutions such as aCSF (which according to our measurements contain anywhere between 4 nm and 50 nm total copper) for prolonged periods of time, one might expect a time-dependent partial depletion of copper to occur. As noted earlier, PrPC contains several copper binding sites whose affinities vary widely from the micromolar to the femtomolar range (Jackson et al. 2001). In media or perfusing solutions containing sub-physiological concentrations of copper (e.g. in the tens of nanomolar, compare with published CSF concentrations noted above), it is conceivable that the lower affinity copper-binding sites on PrPC will be depleted of this metal, potentially significantly altering the physiology or even the survival of the neuron. On the other hand it is important to note that prolonged exposure of neurons to higher levels of copper is also toxic due to the generation of free radicals (Simpson et al. 1988). This may explain findings from Sassoon et al. (2004) showing that prolonged exposure of neurons to 10 μm copper exacerbated the effects of a toxic PrPC fragment.

In our experiments, appearance of persistent currents upon application of BCS occurred with a lag of approximately 2 min, consistent with a mechanism whereby PrPC is initially loaded with copper, and BCS (which binds copper with attomolar affinity) slowly chelates copper ions that unbind from PrPC. This suggests that copper ions are likely to be bound to PrPC (at least at the higher affinity sites) even when bathed in aCSF, either from trace copper in the solutions or from copper ions released by the neurons themselves, or because copper ions may be leaching from the experimental apparatus. Altogether, these considerations suggest that maintenance of consistent copper homeostasis during experiments is critical. By extension, it is conceivable that some of the variability in the extensive published literature pertaining to NMDA receptor physiology could be attributed to variable trace copper levels in media and perfusing fluids. We would therefore argue that copper ions might have to be supplied exogenously to maintain consistent concentrations, similar to what is done routinely with magnesium.

It should be noted that copper ions not only modulate NMDA receptors, but also a variety of other types of ion channels. For example, AMPA receptors in rat cortical neurons are blocked with IC50 values of around 5 μm, and at saturating levels of copper (i.e. 30 μm) agonist affinity for these receptors is reduced (Weiser & Wienrich, 1996). Copper ions also block T-type calcium channels and high voltage activated calcium channels with affinities ranging from ∼1 to 30 μm depending on calcium channel subtype (Jeong et al. 2003; Lu et al. 2009), and they modulate ENaC channels in the submicromolar range (Lu et al. 2009). These blocking affinities are similar to those observed in our own experiments on NMDA receptors in the absence of PrPC, which fits with the idea that voltage gated cation channels may show common architectural features in their pore forming regions – the presumed site of action of copper. In the nanomolar range, copper ions have been shown to enhance the firing of olfactory epithelium neurons, whereas concentrations above 1 μm were inhibitory (Aedo et al. 2007). These potent effects of copper on ion channel physiology and neuronal activity further underscore the need to clamp copper concentrations at a physiological level during in vitro studies of the nervous system.

Aβ-mediated regulation of NMDA receptor function

NMDA receptors have been implicated in Alzheimer's disease (AD), and the NMDA receptor blocker memantine is Food and Drug Administration (FDA) approved for the treatment of this disorder (Kalia et al. 2008; Kotermanski & Johnson, 2009; Di Stefano et al. 2011). One of the key aspects of AD is an overproduction of toxic Aβ1-42 peptides in the brain. Application of Aβ1-42 oligomers to cultured neurons has been shown to trigger spontaneous NMDA currents and longer term exposure of neurons to Aβ1-42 leads to NMDA receptor internalization (Snyder et al. 2005; Texido et al. 2011). It was recently reported by Lauren et al. (2009) that Aβ1-42 oligomers can interact with PrPC and that the toxic effect of Aβ1-42 is dependent on these interactions. Furthermore, Aβ1-42 is a high affinity copper binding peptide (Atwood et al. 2000), with values similar to those reported for BCS. Altogether, these considerations suggest a possible link between Aβ1-42, copper, PrPC and NMDA receptors. Indeed, as we have shown recently (You et al. 2012), application of nanomolar concentrations of Aβ1-42 oligomers, or micromolar concentrations of Aβ1-42 monomers, produced effects on NMDA receptors that were indistinguishable from those of BCS or removal of PrPC (either acutely via GPI anchor cleavage or by gene knockout). Furthermore, recordings from neurons cultured from hemizygous 5XFAD mice, a mouse model of Alzheimer's disease where large amounts of Aβ1-42 are produced within neurons (Oakley et al. 2006), revealed more slowly desensitizing NMDA currents compared to wild-type littermates, again with characteristics very similar to those observed with BCS treatment or PrPC ablation. Like BCS, application of Aβ1-42 to cultured neurons causes cell death that could be prevented either with NMDA receptor antagonists, or by addition of excess copper (You et al. 2012). Together these data suggest that an interplay between copper, Aβ1-42, PrPC and NMDA receptors may play a pivotal role in AD pathogenesis. We envision that high affinity binding of oligomers directly to PrPC mediates an allosteric inhibition of NMDA receptor desensitization (similar to the absence of PrPC), whereas higher concentrations of monomer might act simply by copper chelation (similar to BCS). Therefore, Aβ peptides are likely to mediate their deleterious effects by two related, but distinct mechanisms: one dependent on PrPC where oligomers interfere with normal physiological function (i.e. regulation of NMDA receptor kinetics) of this ubiquitous cuproprotein, and the second by more directly interfering with the ability of copper to modulate these receptors. Such a dual mode of Aβ action would elegantly reconcile the ongoing controversy surrounding the requirement of PrPC for the observed pathological effects of this peptide (Lauren et al. 2009; Calella et al. 2010; Gimbel et al. 2010; Kessels et al. 2010; Freir et al. 2011) (see below).

If elimination of PrPC mediates similar electrophysiological consequences compared to those of Aβ1-42, why do PrP null mice not exhibit synaptic and neuronal degeneration as in AD? There are at least two key differences between the absence of PrPC and the accumulation of Aβ1-42. First, a slowing of NMDA receptor desensitization only matters under conditions where there is prolonged exposure to (i.e. excess) glutamate. Indeed, Aβ1-42 is known to reduce glutamate reuptake in hippocampal neurons (Li et al. 2009), which would lead to increased glutamate concentrations and toxic persistent currents that would not be observed in PrP null mice. Second, as noted earlier, NMDA receptors can be blocked by copper independently of PrPC. Given that higher copper concentrations can compensate for the absence of PrPC by restoring NMDA receptor kinetics to near normalcy, it is possible that in the PrP null brain, compensatory mechanisms are able to maintain free copper in key locations (e.g. in the synaptic cleft) at higher levels to mitigate the adverse influence of a PrPC deficit on this key receptor.

It is important to note that the involvement of PrPC in the toxic effects of Aβ1-42 have become somewhat controversial. In support of the finding of Lauren and colleagues, Barry et al. (2011) reported that suppression of in vivo long-term potentiation by Aβ1-42 oligomers was dependent on PrPC. Freir et al. (2011) reported that Aβ1-42 isolated from human AD brains could alter neuronal plasticity in a manner dependent on PrPC. Finally, Gimbel et al. (2010) reported that deletion of PrPC could abrogate the memory impairment seen in a mouse AD model. However, several other groups were unable to confirm a PrPC dependence of the effects of Aβ1-42 on synaptic physiology (Calella et al. 2010; Kessels et al. 2010). It is possible that different levels of copper and glycine used in these various studies lie at the root of these discordant findings, again stressing the need to maintain copper at a known fixed concentration in neurophysiological studies involving NMDA receptors.

Another perplexing question that arises is, why is PrPSc so toxic to neurons? If complete absence of PrPC, by our arguments a very important NMDA receptor regulator, results in minimal neuronal toxicity (perhaps because ambient copper levels are adjusted accordingly), what is the mechanism of such extreme toxicity of some species of scrapie? The answer is not known, but it is conceivable that prion protein may assume many different conformation states, with PrPSc being the most nefarious. In this conformation, it is possible that the modulatory influence on NMDA receptors is so disadvantageous, potentially inducing high persistent currents that cannot be rescued by any compensatory mechanisms that are invoked in tissue merely lacking PrP. As a result of neurons being locked into a ‘chronic excitotoxic state’, massive degeneration occurs, which has been shown to be at least partially NMDA receptor dependent (Muller et al. 1993; Schroder et al. 1998). Clearly, this hypothesis will have to be tested experimentally.

Concluding remarks

Copper ions appear to mediate potent regulation of NMDA receptors akin to what has been described for zinc nearly two decades ago. Dysregulation of copper homeostasis near the NMDA receptor can trigger aberrant, slowly desensitizing currents that result in toxic calcium overload in the presence of sustained glutamate levels. This in turn may contribute to the pathology of Alzheimer's disease. Copper dysregulation of NMDA receptors may also play a role in other neurological disorders. For example, in Menkes disease, a mutation in the copper ATPase mediates NMDA receptor-dependent neurodegeneration (Schlief et al. 2005, 2006; Schlief & Gitlin, 2006). Furthermore, both α-synuclein and huntingtin are known to be copper binding proteins (Fox et al. 2007; Davies et al. 2010), and both have been associated with NMDA receptor mediated neuronal toxicity (Fan & Raymond, 2007; Adamczyk et al. 2009). This raises the possibility of copper modulation of NMDA receptors (perhaps via PrPC) as a unifying theme in many neurodegenerative disorders.

Acknowledgments

G.W.Z. and P.K.S. are Alberta Innovates–Health Solutions (AIHS) Scientists and hold Canada Research Chairs. Work from the authors’ laboratories discussed in this review was funded by the Canadian Institutes of Health Research, The Alberta Prion Research Institute, and PrioNet Canada. H.Y. is an AIHS postdoctoral fellow.

References

  1. Aarts M, et al. Treatment of ischemic brain damage by perturbing NMDA receptor- PSD-95 protein interactions. Science. 2002;298:846–850. doi: 10.1126/science.1072873. [DOI] [PubMed] [Google Scholar]
  2. Adamczyk A, Czapski GA, Kazmierczak A, Strosznajder JB. Effect of N-methyl-D-aspartate (NMDA) receptor antagonists on α-synuclein-evoked neuronal nitric oxide synthase activation in the rat brain. Pharmacol Rep. 2009;61:1078–1085. doi: 10.1016/s1734-1140(09)70170-7. [DOI] [PubMed] [Google Scholar]
  3. Aedo F, Delgado R, Wolff D, Vergara C. Copper and zinc as modulators of neuronal excitability in a physiologically significant concentration range. Neurochem Int. 2007;50:591–600. doi: 10.1016/j.neuint.2006.12.001. [DOI] [PubMed] [Google Scholar]
  4. Agarwal RP, Henkin RI. Zinc and copper in human cerebrospinal fluid. Biol Trace Element Res. 1982;4:117–124. doi: 10.1007/BF02783252. [DOI] [PubMed] [Google Scholar]
  5. Aguzzi A, Heikenwalder M. Pathogenesis of prion diseases: current status and future outlook. Nat Rev Microbiol. 2006;4:765–775. doi: 10.1038/nrmicro1492. [DOI] [PubMed] [Google Scholar]
  6. Aguzzi A, Sigurdson C, Heikenwaelder M. Molecular mechanisms of prion pathogenesis. Annu Rev Pathol. 2008;3:11–40. doi: 10.1146/annurev.pathmechdis.3.121806.154326. [DOI] [PubMed] [Google Scholar]
  7. Al-Shatti N, Lappin AG, Sykes AG. Influence of coordination number on copper(I)-Copper(II) redox interconversions. 2. Fe(CN)64- reduction of a sterically constrained bis(substituted phenanthroline) complex of copper(II) in aqueous solution. Inorg Chem. 1981;20:1466–1469. [Google Scholar]
  8. Alier K, Li Z, Mactavish D, Westaway D, Jhamandas JH. Ionic mechanisms of action of prion protein fragment PrP(106–126) in rat basal forebrain neurons. J Neurosci Res. 2010;88:2217–2227. doi: 10.1002/jnr.22372. [DOI] [PubMed] [Google Scholar]
  9. Asztely F, Erdemli G, Kullmann DM. Extrasynaptic glutamate spillover in the hippocampus: dependence on temperature and the role of active glutamate uptake. Neuron. 1997;18:281–293. doi: 10.1016/s0896-6273(00)80268-8. [DOI] [PubMed] [Google Scholar]
  10. Atwood CS, et al. Characterization of copper interactions with Alzheimer amyloid beta peptides: identification of an attomolar-affinity copper binding site on amyloid β1–42. J Neurochem. 2000;75:1219–1233. doi: 10.1046/j.1471-4159.2000.0751219.x. [DOI] [PubMed] [Google Scholar]
  11. Barry AE, et al. Alzheimer's disease brain-derived amyloid-β-mediated inhibition of LTP in vivo is prevented by immunotargeting cellular prion protein. J Neurosci. 2011;31:7259–7263. doi: 10.1523/JNEUROSCI.6500-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Beraldo FH, et al. Metabotropic glutamate receptors transduce signals for neurite outgrowth after binding of the prion protein to laminin γ1 chain. FASEB J. 2011;25:265–279. doi: 10.1096/fj.10-161653. [DOI] [PubMed] [Google Scholar]
  13. Beraldo FH, et al. Role of α7 nicotinic acetylcholine receptor in calcium signaling induced by prion protein interaction with stress-inducible protein 1. J Biol Chem. 2010;285:36542–36550. doi: 10.1074/jbc.M110.157263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bremer J, et al. Axonal prion protein is required for peripheral myelin maintenance. Nat Neurosci. 2010;13:310–318. doi: 10.1038/nn.2483. [DOI] [PubMed] [Google Scholar]
  15. Brown DR, et al. The cellular prion protein binds copper in vivo. Nature. 1997;390:684–687. doi: 10.1038/37783. [DOI] [PubMed] [Google Scholar]
  16. Brown DR, Sassoon J. Copper-dependent functions for the prion protein. Mol Biotechnol. 2002;22:165–178. doi: 10.1385/MB:22:2:165. [DOI] [PubMed] [Google Scholar]
  17. Bueler H, et al. Normal development and behaviour of mice lacking the neuronal cell-surface PrP protein. Nature. 1992;356:577–582. doi: 10.1038/356577a0. [DOI] [PubMed] [Google Scholar]
  18. Burns CS, et al. Copper coordination in the full-length, recombinant prion protein. Biochemistry. 2003;42:6794–6803. doi: 10.1021/bi027138+. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Calella AM, et al. Prion protein and Aβ-related synaptic toxicity impairment. EMBO Mol Med. 2010;2:306–314. doi: 10.1002/emmm.201000082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Caughey B, Baron GS, Chesebro B, Jeffrey M. Getting a grip on prions: oligomers, amyloids, and pathological membrane interactions. Annu Rev Biochem. 2009;78:177–204. doi: 10.1146/annurev.biochem.78.082907.145410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Chatterton JE, et al. Excitatory glycine receptors containing the NR3 family of NMDA receptor subunits. Nature. 2002;415:793–798. doi: 10.1038/nature715. [DOI] [PubMed] [Google Scholar]
  22. Chen L, Muhlhauser M, Yang CR. Glycine tranporter-1 blockade potentiates NMDA-mediated responses in rat prefrontal cortical neurons in vitro and in vivo. J Neurophysiol. 2003;89:691–703. doi: 10.1152/jn.00680.2002. [DOI] [PubMed] [Google Scholar]
  23. Chen PE, et al. Modulation of glycine potency in rat recombinant NMDA receptors containing chimeric NR2A/2D subunits expressed in Xenopus laevis oocytes. J Physiol. 2008;586:227–245. doi: 10.1113/jphysiol.2007.143172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Choi YB, Lipton SA. Identification and mechanism of action of two histidine residues underlying high-affinity Zn2+ inhibition of the NMDA receptor. Neuron. 1999;23:171–180. doi: 10.1016/s0896-6273(00)80763-1. [DOI] [PubMed] [Google Scholar]
  25. Collinge J, et al. Prion protein is necessary for normal synaptic function. Nature. 1994;370:295–297. doi: 10.1038/370295a0. [DOI] [PubMed] [Google Scholar]
  26. Collingridge G. Synaptic plasticity. The role of NMDA receptors in learning and memory. Nature. 1987;330:604–605. doi: 10.1038/330604a0. [DOI] [PubMed] [Google Scholar]
  27. Collingridge GL, Olsen RW, Peters J, Spedding M. A nomenclature for ligand-gated ion channels. Neuropharmacology. 2009;56:2–5. doi: 10.1016/j.neuropharm.2008.06.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Criado JR, et al. Mice devoid of prion protein have cognitive deficits that are rescued by reconstitution of PrP in neurons. Neurobiol Dis. 2005;19:255–265. doi: 10.1016/j.nbd.2005.01.001. [DOI] [PubMed] [Google Scholar]
  29. Cull-Candy SG, Leszkiewicz DN. Role of distinct NMDA receptor subtypes at central synapses. Sci STKE. 2004;2004:re16. doi: 10.1126/stke.2552004re16. [DOI] [PubMed] [Google Scholar]
  30. Curtis J, Errington M, Bliss T, Voss K, MacLeod N. Age-dependent loss of PTP and LTP in the hippocampus of PrP-null mice. Neurobiol Dis. 2003;13:55–62. doi: 10.1016/s0969-9961(03)00017-2. [DOI] [PubMed] [Google Scholar]
  31. Davies P, et al. The synucleins are a family of redox-active copper binding proteins. Biochemistry. 2010;50:37–47. doi: 10.1021/bi101582p. [DOI] [PubMed] [Google Scholar]
  32. Di Stefano A, Iannitelli A, Laserra S, Sozio P. Drug delivery strategies for Alzheimer's disease treatment. Expert Opin Drug Deliv. 2011;8:581–603. doi: 10.1517/17425247.2011.561311. [DOI] [PubMed] [Google Scholar]
  33. Diamond JS. Neuronal glutamate transporters limit activation of NMDA receptors by neurotransmitter spillover on CA1 pyramidal cells. J Neurosci. 2001;21:8328–8338. doi: 10.1523/JNEUROSCI.21-21-08328.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Erreger K, et al. Subunit-specific agonist activity at NR2A-, NR2B-, NR2C-, and NR2D-containing N-methyl-D-aspartate glutamate receptors. Mol Pharmacol. 2007;72:907–920. doi: 10.1124/mol.107.037333. [DOI] [PubMed] [Google Scholar]
  35. Fan MM, Raymond LA. N-methyl-D-aspartate (NMDA) receptor function and excitotoxicity in Huntington's disease. Prog Neurobiol. 2007;81:272–293. doi: 10.1016/j.pneurobio.2006.11.003. [DOI] [PubMed] [Google Scholar]
  36. Fox JH, et al. Mechanisms of copper ion mediated Huntington's disease progression. PLoS One. 2007;2:e334. doi: 10.1371/journal.pone.0000334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Freir DB, et al. Interaction between prion protein and toxic amyloid β assemblies can be therapeutically targeted at multiple sites. Nat Commun. 2011;2:336. doi: 10.1038/ncomms1341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Gadotti VM, Bonfield SP, Zamponi GW. Depressive-like behaviour of mice lacking cellular prion protein. Behav Brain Res. 2012;227:319–323. doi: 10.1016/j.bbr.2011.03.012. [DOI] [PubMed] [Google Scholar]
  39. Gadotti VM, Zamponi GW. Cellular prion protein protects from inflammatory and neuropathic pain. Mol Pain. 2011;7:59. doi: 10.1186/1744-8069-7-59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Gajdusek DC. Unconventional viruses and the origin and disappearance of kuru. Science. 1977;197:943–960. doi: 10.1126/science.142303. [DOI] [PubMed] [Google Scholar]
  41. Gimbel DA, et al. Memory impairment in transgenic Alzheimer mice requires cellular prion protein. J Neurosci. 2010;30:6367–6374. doi: 10.1523/JNEUROSCI.0395-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Griffin JK, Terry LA, Jackman R, Yousefi M, Cashman NR. Decreased cell surface prion protein in mouse models of prion disease. Neuroreport. 2007;18:1–6. doi: 10.1097/01.wnr.0000239967.06438.21. [DOI] [PubMed] [Google Scholar]
  43. Herms JW, Tings T, Dunker S, Kretzschmar HA. Prion protein affects Ca2+-activated K+ currents in cerebellar purkinje cells. Neurobiol Dis. 2001;8:324–330. doi: 10.1006/nbdi.2000.0369. [DOI] [PubMed] [Google Scholar]
  44. Herron CE, Lester RA, Coan EJ, Collingridge GL. Frequency-dependent involvement of NMDA receptors in the hippocampus: a novel synaptic mechanism. Nature. 1986;322:265–268. doi: 10.1038/322265a0. [DOI] [PubMed] [Google Scholar]
  45. Hill AF, et al. The same prion strain causes vCJD and BSE. Nature. 1997;389:448–50. 526. doi: 10.1038/38925. [DOI] [PubMed] [Google Scholar]
  46. Hu W, et al. Pharmacological prion protein silencing accelerates central nervous system autoimmune disease via T cell receptor signalling. Brain. 2010;133:375–388. doi: 10.1093/brain/awp298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Huang T, et al. PrP(C) interacts with potassium channel tetramerization domain containing 1 (KCTD1) protein through the PrP(51–136) region containing octapeptide repeats. Biochem Biophys Res Commun. 2012;417:182–186. doi: 10.1016/j.bbrc.2011.11.081. [DOI] [PubMed] [Google Scholar]
  48. Jackson GS, et al. Location and properties of metal-binding sites on the human prion protein. Proc Natl Acad Sci U S A. 2001;98:8531–8535. doi: 10.1073/pnas.151038498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Jeong SW, Park BG, Park JY, Lee JW, Lee JH. Divalent metals differentially block cloned T-type calcium channels. Neuroreport. 2003;14:1537–1540. doi: 10.1097/00001756-200308060-00028. [DOI] [PubMed] [Google Scholar]
  50. Johnston AR, Fraser JR, Jeffrey M, MacLeod N. Synaptic plasticity in the CA1 area of the hippocampus of scrapie-infected mice. Neurobiol Dis. 1998;5:188–195. doi: 10.1006/nbdi.1998.0194. [DOI] [PubMed] [Google Scholar]
  51. Kalia LV, Kalia SK, Salter MW. NMDA receptors in clinical neurology: excitatory times ahead. Lancet Neurol. 2008;7:742–755. doi: 10.1016/S1474-4422(08)70165-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Kanabrocki EL, Case LF, Miller EB, Kaplan E, Oester YT. A study of human cerebrospinal fluid: copper and manganese. J Nucl Med. 1964;5:643–648. [PubMed] [Google Scholar]
  53. Kessels HW, Nguyen LN, Nabavi S, Malinow R. The prion protein as a receptor for amyloid-β. Nature. 2010;466:E3–5. doi: 10.1038/nature09217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Khosravani H, et al. Prion protein attenuates excitotoxicity by inhibiting NMDA receptors. J Cell Biol. 2008;181:551–565. doi: 10.1083/jcb.200711002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Kingsbury DT, Kasper KC, Stites DP, Watson JD, Hogan RN, Prusiner SB. Genetic control of scrapie and Creutzfeldt-Jakob disease in mice. J Immunol. 1983;131:491–496. [PubMed] [Google Scholar]
  56. Korte S, Vassallo N, Kramer ML, Kretzschmar HA, Herms J. Modulation of L-type voltage-gated calcium channels by recombinant prion protein. J Neurochem. 2003;87:1037–1042. doi: 10.1046/j.1471-4159.2003.02080.x. [DOI] [PubMed] [Google Scholar]
  57. Kotermanski SE, Johnson JW. Mg2+ imparts NMDA receptor subtype selectivity to the Alzheimer's drug memantine. J Neurosci. 2009;29:2774–2779. doi: 10.1523/JNEUROSCI.3703-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Kupper J, Ascher P, Neyton J. Probing the pore region of recombinant N-methyl-D-aspartate channels using external and internal magnesium block. Proc Natl Acad Sci U S A. 1996;93:8648–8653. doi: 10.1073/pnas.93.16.8648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Lauren J, Gimbel DA, Nygaard HB, Gilbert JW, Strittmatter SM. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-β oligomers. Nature. 2009;457:1128–1132. doi: 10.1038/nature07761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Lee FJ, et al. Dual regulation of NMDA receptor functions by direct protein-protein interactions with the dopamine D1 receptor. Cell. 2002;111:219–230. doi: 10.1016/s0092-8674(02)00962-5. [DOI] [PubMed] [Google Scholar]
  61. Li S, Hong S, Shepardson NE, Walsh DM, Shankar GM, Selkoe D. Soluble oligomers of amyloid beta protein facilitate hippocampal long-term depression by disrupting neuronal glutamate uptake. Neuron. 2009;62:788–801. doi: 10.1016/j.neuron.2009.05.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Linden R, Martins VR, Prado MA, Cammarota M, Izquierdo I, Brentani RR. Physiology of the prion protein. Physiol Rev. 2008;88:673–728. doi: 10.1152/physrev.00007.2007. [DOI] [PubMed] [Google Scholar]
  63. Lipton SA, Rosenberg PA. Excitatory amino acids as a final common pathway for neurologic disorders. N Engl J Med. 1994;330:613–622. doi: 10.1056/NEJM199403033300907. [DOI] [PubMed] [Google Scholar]
  64. Lipton SA. Paradigm shift in neuroprotection by NMDA receptor blockade: memantine and beyond. Nat Rev Drug Discov. 2006;5:160–170. doi: 10.1038/nrd1958. [DOI] [PubMed] [Google Scholar]
  65. Liu XJ, et al. Treatment of inflammatory and neuropathic pain by uncoupling Src from the NMDA receptor complex. Nat Med. 2008;14:1325–1332. doi: 10.1038/nm.1883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Liu Y, et al. NMDA receptor subunits have differential roles in mediating excitotoxic neuronal death both in vitro and in vivo. J Neurosci. 2007;27:2846–2857. doi: 10.1523/JNEUROSCI.0116-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Loftis JM, Janowsky A. The N-methyl-D-aspartate receptor subunit NR2B: localization, functional properties, regulation, and clinical implications. Pharmacol Ther. 2003;97:55–85. doi: 10.1016/s0163-7258(02)00302-9. [DOI] [PubMed] [Google Scholar]
  68. Low CM, Wee KS. New insights into the not-so-new NR3 subunits of N-methyl-D-aspartate receptor: localization, structure, and function. Mol Pharmacol. 2010;78:1–11. doi: 10.1124/mol.110.064006. [DOI] [PubMed] [Google Scholar]
  69. Lozovaya NA, Kopanitsa MV, Boychuk YA, Krishtal OA. Enhancement of glutamate release uncovers spillover-mediated transmission by N-methyl-D-aspartate receptors in the rat hippocampus. Neuroscience. 1999;91:1321–1330. doi: 10.1016/s0306-4522(98)00638-1. [DOI] [PubMed] [Google Scholar]
  70. Lu L, et al. Effects of copper on T-type Ca2+ channels in mouse spermatogenic cells. J Membr Biol. 2009;227:87–94. doi: 10.1007/s00232-008-9148-y. [DOI] [PubMed] [Google Scholar]
  71. MacDonald JF, Jackson MF, Beazely MA. Hippocampal long-term synaptic plasticity and signal amplification of NMDA receptors. Crit Rev Neurobiol. 2006;18:71–84. doi: 10.1615/critrevneurobiol.v18.i1-2.80. [DOI] [PubMed] [Google Scholar]
  72. Mallucci GR, et al. Inherited prion disease with an alanine to valine mutation at codon 117 in the prion protein gene. Brain. 1999;122:1823–1837. doi: 10.1093/brain/122.10.1823. [DOI] [PubMed] [Google Scholar]
  73. Martin RP, Mosoni L, Sarkar B. Ternary coordination complexes between glycine, copper (II), and glycine peptides in aqueous solution. J Biol Chem. 1971;246:5944–5951. [PubMed] [Google Scholar]
  74. Mayer ML, Westbrook GL, Guthrie PB. Voltage-dependent block by Mg2+ of NMDA responses in spinal cord neurones. Nature. 1984;309:261–263. doi: 10.1038/309261a0. [DOI] [PubMed] [Google Scholar]
  75. Mayer ML, Vyklicky LJ, Clements J. Regulation of NMDA receptor desensitization in mouse hippocampal neurons by glycine. Nature. 1989;338:425–427. doi: 10.1038/338425a0. [DOI] [PubMed] [Google Scholar]
  76. McBain CJ, Mayer ML. N-methyl-D-aspartic acid receptor structure and function. Physiol Rev. 1994;74:723–760. doi: 10.1152/physrev.1994.74.3.723. [DOI] [PubMed] [Google Scholar]
  77. Millhauser GL. Copper and the prion protein: methods, structures, function, and disease. Annu Rev Phys Chem. 2007;58:299–320. doi: 10.1146/annurev.physchem.58.032806.104657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Mori H, Masaki H, Yamakura T, Mishina M. Identification by mutagenesis of a Mg2+-block site of the NMDA receptor channel. Nature. 1992;358:673–675. doi: 10.1038/358673a0. [DOI] [PubMed] [Google Scholar]
  79. Muller WE, et al. Cytoprotective effect of NMDA receptor antagonists on prion protein (PrionSc)-induced toxicity in rat cortical cell cultures. Eur J Pharmacol. 1993;246:261–267. doi: 10.1016/0922-4106(93)90040-g. [DOI] [PubMed] [Google Scholar]
  80. Niciu MJ, Ma XM, El Meskini R, Ronnett GV, Mains RE, Eipper BA. Developmental changes in the expression of ATP7A during a critical period in postnatal neurodevelopment. Neuroscience. 2006;139:947–964. doi: 10.1016/j.neuroscience.2006.01.044. [DOI] [PubMed] [Google Scholar]
  81. Nishida N, et al. Prion protein is necessary for latent learning and long-term memory retention. Cell Mol Neurobiol. 1997;17:537–545. doi: 10.1023/a:1026315006619. [DOI] [PubMed] [Google Scholar]
  82. Nowak L, Bregestovski P, Ascher P, Herbet A, Prochiantz A. Magnesium gates glutamate-activated channels in mouse central neurones. Nature. 1984;307:462–465. doi: 10.1038/307462a0. [DOI] [PubMed] [Google Scholar]
  83. Oakley H, et al. Intraneuronal β-amyloid aggregates, neurodegeneration, and neuron loss in transgenic mice with five familial Alzheimer's disease mutations: potential factors in amyloid plaque formation. J Neurosci. 2006;26:10129–10140. doi: 10.1523/JNEUROSCI.1202-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Olivares D, et al. N-Methyl D-aspartate (NMDA) receptor antagonists and memantine treatment for Alzheimer's disease, vascular dementia and Parkinson's disease. Curr Alzheimer Res. 2011 doi: 10.2174/156720512801322564. (in press) [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Palmer MS, Dryden AJ, Hughes JT, Collinge J. Homozygous prion protein genotype predisposes to sporadic Creutzfeldt-Jakob disease. Nature. 1991;352:340–342. doi: 10.1038/352340a0. [DOI] [PubMed] [Google Scholar]
  86. Paoletti P. Molecular basis of NMDA receptor functional diversity. Eur J Neurosci. 2011;33:1351–1365. doi: 10.1111/j.1460-9568.2011.07628.x. [DOI] [PubMed] [Google Scholar]
  87. Paoletti P, Neyton J. NMDA receptor subunits: function and pharmacology. Curr Opin Pharmacol. 2007;7:39–47. doi: 10.1016/j.coph.2006.08.011. [DOI] [PubMed] [Google Scholar]
  88. Peters S, Koh J, Choi DW. Zinc selectively blocks the action of N-methyl-D-aspartate on cortical neurons. Science. 1987;236:589–593. doi: 10.1126/science.2883728. [DOI] [PubMed] [Google Scholar]
  89. Peterson RE, Bollier ME. Spectrophotometric determination of serum copper with biscyclohexanoneoxalyldihydrazone. Anal Chem. 1955;27:1195–1197. [Google Scholar]
  90. Piña-Crespo JC, et al. Excitatory glycine responses of CNS myelin mediated by NR1/NR3 ‘NMDA’ receptor subunits. J Neurosci. 2010;30:11501–11505. doi: 10.1523/JNEUROSCI.1593-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Prusiner SB. Prion diseases and the BSE crisis. Science. 1997;278:245–251. doi: 10.1126/science.278.5336.245. [DOI] [PubMed] [Google Scholar]
  92. Rachline J, Perin-Dureau F, Le Goff A, Neyton J, Paoletti P. The micromolar zinc-binding domain on the NMDA receptor subunit NR2B. J Neurosci. 2005;25:308–317. doi: 10.1523/JNEUROSCI.3967-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Rangel A, Burgaya F, Gavin R, Soriano E, Aguzzi A, Del Rio JA. Enhanced susceptibility of Prnp-deficient mice to kainate-induced seizures, neuronal apoptosis, and death: Role of AMPA/kainate receptors. J Neurosci Res. 2007;85:2741–2755. doi: 10.1002/jnr.21215. [DOI] [PubMed] [Google Scholar]
  94. Rangel A, et al. Regulation of GABAA and glutamate receptor expression, synaptic facilitation and long-term potentiation in the hippocampus of prion mutant mice. PLoS One. 2009;4:e7592. doi: 10.1371/journal.pone.0007592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Ratte S, et al. Threshold for epileptiform activity is elevated in prion knockout mice. Neuroscience. 2011;179:56–61. doi: 10.1016/j.neuroscience.2011.01.053. [DOI] [PubMed] [Google Scholar]
  96. Rauner C, Kohr G. Triheteromeric NR1/NR2A/NR2B receptors constitute the major N-methyl-D-aspartate receptor population in adult hippocampal synapses. J Biol Chem. 2011;286:7558–7566. doi: 10.1074/jbc.M110.182600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Riesner D. Biochemistry and structure of PrPC and PrPSc. Br Med Bull. 2003;66:21–33. doi: 10.1093/bmb/66.1.21. [DOI] [PubMed] [Google Scholar]
  98. Sailer A, Bueler H, Fischer M, Aguzzi A, Weissmann C. No propagation of prions in mice devoid of PrP. Cell. 1994;77:967–968. doi: 10.1016/0092-8674(94)90436-7. [DOI] [PubMed] [Google Scholar]
  99. Salter MW. Cellular signalling pathways of spinal pain neuroplasticity as targets for analgesic development. Curr Top Med Chem. 2005;5:557–567. doi: 10.2174/1568026054367638. [DOI] [PubMed] [Google Scholar]
  100. Sanhueza M, et al. Role of the CaMKII/NMDA receptor complex in the maintenance of synaptic strength. J Neurosci. 2011;31:9170–9178. doi: 10.1523/JNEUROSCI.1250-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Sassoon J, Daniels M, Brown DR. Astrocytic regulation of NMDA receptor subunit composition modulates the toxicity of prion peptide PrP106–126. Mol Cell Neurosci. 2004;25:181–191. doi: 10.1016/j.mcn.2003.09.019. [DOI] [PubMed] [Google Scholar]
  102. Sayre LM. Alzheimer's precursor protein and the use of bathocuproine for determining reduction of copper(II) Science. 1996;274:1933–1934. doi: 10.1126/science.274.5294.1933. [DOI] [PubMed] [Google Scholar]
  103. Schlief ML, Craig AM, Gitlin JD. NMDA receptor activation mediates copper homeostasis in hippocampal neurons. J Neurosci. 2005;25:239–246. doi: 10.1523/JNEUROSCI.3699-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Schlief ML, Gitlin JD. Copper homeostasis in the CNS: a novel link between the NMDA receptor and copper homeostasis in the hippocampus. Mol Neurobiol. 2006;33:81–90. doi: 10.1385/MN:33:2:81. [DOI] [PubMed] [Google Scholar]
  105. Schlief ML, West T, Craig AM, Holtzman DM, Gitlin JD. Role of the Menkes copper-transporting ATPase in NMDA receptor-mediated neuronal toxicity. Proc Natl Acad Sci U S A. 2006;103:14919–14924. doi: 10.1073/pnas.0605390103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Schroder HC, Perovic S, Kavsan V, Ushijima H, Muller WE. Mechanisms of prionSc- and HIV-1 gp120 induced neuronal cell death. Neurotoxicology. 1998;19:683–688. [PubMed] [Google Scholar]
  107. Shleper M, Kartvelishvily E, Wolosker H. D-Serine is the dominant endogenous coagonist for NMDA receptor neurotoxicity in organotypic hippocampal slices. J Neurosci. 2005;25:9413–9417. doi: 10.1523/JNEUROSCI.3190-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Simpson JA, Cheeseman KH, Smith SE, Dean RT. Free-radical generation by copper ions and hydrogen peroxide. Stimulation by Hepes buffer. Biochem J. 1988;254:519–523. doi: 10.1042/bj2540519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Snyder EM, et al. Regulation of NMDA receptor trafficking by amyloid-beta. Nat Neurosci. 2005;8:1051–1058. doi: 10.1038/nn1503. [DOI] [PubMed] [Google Scholar]
  110. Steele AD, Lindquist S, Aguzzi A. The prion protein knockout mouse: a phenotype under challenge. Prion. 2007;1:83–93. doi: 10.4161/pri.1.2.4346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Stotz SC, Jarvis SE, Zamponi GW. Functional roles of cytoplasmic loops and pore lining transmembrane helices in the voltage-dependent inactivation of HVA calcium channels. J Physiol. 2004;554:263–273. doi: 10.1113/jphysiol.2003.047068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Supplisson S, Roux MJ. Why glycine transporters have different stoichiometries. FEBS Lett. 2002;529:93–101. doi: 10.1016/s0014-5793(02)03251-9. [DOI] [PubMed] [Google Scholar]
  113. Texido L, Martin-Satue M, Alberdi E, Solsona C, Matute C. Amyloid beta peptide oligomers directly activate NMDA receptors. Cell Calcium. 2011;49:184–190. doi: 10.1016/j.ceca.2011.02.001. [DOI] [PubMed] [Google Scholar]
  114. Thomas CG, Miller AJ, Westbrook GL. Synaptic and extrasynaptic NMDA receptor NR2 subunits in cultured hippocampal neurons. J Neurophysiol. 2006;95:1727–1734. doi: 10.1152/jn.00771.2005. [DOI] [PubMed] [Google Scholar]
  115. Tingley WG, Roche KW, Thompson AK, Huganir RL. Regulation of NMDA receptor phosphorylation by alternative splicing of the C-terminal domain. Nature. 1993;364:70–73. doi: 10.1038/364070a0. [DOI] [PubMed] [Google Scholar]
  116. Trepanier CH, Jackson MF, Macdonald JF. Regulation of NMDA receptors by the tyrosine kinase Fyn. FEBS J. 2012;279:12–19. doi: 10.1111/j.1742-4658.2011.08391.x. [DOI] [PubMed] [Google Scholar]
  117. Tsutsui S, Hahn JN, Johnson TA, Ali Z, Jirik FR. Absence of the cellular prion protein exacerbates and prolongs neuroinflammation in experimental autoimmune encephalomyelitis. Am J Pathol. 2008;173:1029–1041. doi: 10.2353/ajpath.2008.071062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Valenti P, Cozzio A, Nishida N, Wolfer DP, Sakaguchi S, Lipp HP. Similar target, different effects: late-onset ataxia and spatial learning in prion protein-deficient mouse lines. Neurogenetics. 2001;3:173–184. doi: 10.1007/s100480100117. [DOI] [PubMed] [Google Scholar]
  119. Vance KM, Simorowski N, Traynelis SF, Furukawa H. Ligand-specific deactivation time course of GluN1/GluN2D NMDA receptors. Nat Commun. 2011;2:294. doi: 10.1038/ncomms1295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Veldhuis NA, Gaeth AP, Pearson RB, Gabriel K, Camakaris J. The multi-layered regulation of copper translocating P-type ATPases. Biometals. 2009;22:177–190. doi: 10.1007/s10534-008-9183-2. [DOI] [PubMed] [Google Scholar]
  121. Vlachova V, Zemkova H, Vyklicky LJ. Copper modulation of NMDA responses in mouse and rat cultured hippocampal neurons. Eur J Neurosci. 1996;8:2257–2264. doi: 10.1111/j.1460-9568.1996.tb01189.x. [DOI] [PubMed] [Google Scholar]
  122. Vyklicky LJ, Benveniste M, Mayer ML. Modulation of N-methyl-D-aspartic acid receptor desensitization by glycine in mouse cultured hippocampal neurones. J Physiol. 1990;428:313–331. doi: 10.1113/jphysiol.1990.sp018214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Walz R, et al. Increased sensitivity to seizures in mice lacking cellular prion protein. Epilepsia. 1999;40:1679–1682. doi: 10.1111/j.1528-1157.1999.tb01583.x. [DOI] [PubMed] [Google Scholar]
  124. Warren PJ, Earl CJ, Thompson RH. The distribution of copper in human brain. Brain. 1960;83:709–717. doi: 10.1093/brain/83.4.709. [DOI] [PubMed] [Google Scholar]
  125. Weiser T, Wienrich M. The effects of copper ions on glutamate receptors in cultured rat cortical neurons. Brain Res. 1996;742:211–218. doi: 10.1016/s0006-8993(96)01009-8. [DOI] [PubMed] [Google Scholar]
  126. White MD, Farmer M, Mirabile I, Brandner S, Collinge J, Mallucci GR. Single treatment with RNAi against prion protein rescues early neuronal dysfunction and prolongs survival in mice with prion disease. Proc Natl Acad Sci U S A. 2008;105:10238–10243. doi: 10.1073/pnas.0802759105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. White MD, Mallucci GR. RNAi for the treatment of prion disease: a window for intervention in neurodegeneration? CNS Neurol Disord Drug Targets. 2009;8:342–352. doi: 10.2174/187152709789541934. [DOI] [PubMed] [Google Scholar]
  128. Wong BS, et al. Mapping the antigenicity of copper-treated cellular prion protein with the scrapie isoform. Cell Mol Life Sci. 2003;60:1224–1234. doi: 10.1007/s00018-003-3057-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Woolf CJ, Thompson SW. The induction and maintenance of central sensitization is dependent on N-methyl-D-aspartic acid receptor activation; implications for the treatment of post-injury pain hypersensitivity states. Pain. 1991;44:293–299. doi: 10.1016/0304-3959(91)90100-C. [DOI] [PubMed] [Google Scholar]
  130. Xiao Z, Brose J, Schimo S, Ackland SM, La Fontaine S, Wedd AG. Unification of the copper(I) binding affinities of the metallo-chaperones Atx1, Atox1, and related proteins: detection probes and affinity standards. J Biol Chem. 2011;286:11047–11055. doi: 10.1074/jbc.M110.213074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Xu J, et al. Control of excitatory synaptic transmission by C-terminal Src kinase. J Biol Chem. 2008;283:17503–17514. doi: 10.1074/jbc.M800917200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Yang CR, Svensson KA. Allosteric modulation of NMDA receptor via elevation of brain glycine and D-serine: the therapeutic potentials for schizophrenia. Pharmacol Ther. 2008;120:317–332. doi: 10.1016/j.pharmthera.2008.08.004. [DOI] [PubMed] [Google Scholar]
  133. You H, et al. Aβ damages neurons by altering copper-dependent prion protein regulation of NMDA receptors. Proc Natl Acad Sci U S A. 2012;109:1737–1742. [Google Scholar]
  134. Zhang W, Howe JR, Popescu GK. Distinct gating modes determine the biphasic relaxation of NMDA receptor currents. Nat Neurosci. 2008;11:1373–1375. doi: 10.1038/nn.2214. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES