首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The aggregation of amyloids into toxic oligomers is believed to be a key pathogenic event in the onset of Alzheimer''s disease. Peptidomimetic modulators capable of destabilizing the propagation of an extended network of β-sheet fibrils represent a potential intervention strategy. Modifications to amyloid-beta (Aβ) peptides derived from the core domain have afforded inhibitors capable of both antagonizing aggregation and reducing amyloid toxicity. Previous work from our laboratory has shown that peptide backbone amination stabilizes β-sheet-like conformations and precludes β-strand aggregation. Here, we report the synthesis of N-aminated hexapeptides capable of inhibiting the fibrillization of full-length Aβ42. A key feature of our design is N-amino substituents at alternating backbone amides within the aggregation-prone Aβ16–21 sequence. This strategy allows for maintenance of an intact hydrogen-bonding backbone edge as well as side chain moieties important for favorable hydrophobic interactions. An N-amino scan of Aβ16–21 resulted in the identification of peptidomimetics that block Aβ42 fibrilization in several biophysical assays.

Structure-based design of backbone-aminated peptides affords novel β-strand mimics that inhibit amyloid-beta fibrillogenesis.  相似文献   

2.
Protein persulfidation plays a role in redox signaling as an anti-oxidant. Dimers of amyloid β42 (Aβ42), which induces oxidative stress-associated neurotoxicity as a causative agent of Alzheimer''s disease (AD), are minimum units of oligomers in AD pathology. Met35 can be susceptible to persulfidation through its substitution to homoCys residue under the condition of oxidative stress. In order to verify whether persulfidation has an effect in AD, herein we report a chemical approach by synthesizing disulfide dimers of Aβ42 and their evaluation of biochemical properties. A homoCys-disulfide dimer model at position 35 of Aβ42 formed a partial β-sheet structure, but its neurotoxicity was much weaker than that of the corresponding monomer. In contrast, the congener with an alkyl linker generated β-sheet-rich 8–16-mer oligomers with potent neurotoxicity. The length of protofibrils generated from the homoCys-disulfide dimer model was shorter than that of its congener with an alkyl linker. Therefore, the current data do not support the involvement of Aβ42 persulfidation in Alzheimer''s disease.

Our data do not support the Aβ42 persulfidation hypothesis in Alzheimer''s etiology because the neurotoxicity of the homoCys-disulfide-Aβ42 dimer was very weak.  相似文献   

3.
Alzheimer''s disease is linked to the aggregation of the amyloid-β protein (Aβ) of 40 or 42 amino acids. Lipid membranes are known to modulate the rate and mechanisms of the Aβ aggregation. Point mutations in Aβ can alter these rates and mechanisms. In particular, experiments show that F19 mutations influence the aggregation rate, but maintain the fibril structures. Here, we used molecular dynamics simulations to examine the effect of the F19W mutation in the 3Aβ11–40 trimer immersed in DPPC lipid bilayers submerged in aqueous solution. Substituting Phe by its closest (non-polar) aromatic amino acid Trp has a dramatic reduction in binding affinity to the phospholipid membrane (measured with respect to the solvated protein) compared to the wild type: the binding free energy of the protein–DPPC lipid bilayer increases by 40–50 kcal mol−1 over the wild-type. This is accompanied by conformational changes and loss of salt bridges, as well as a more complex free energy surface, all indicative of a more flexible and less stable mutated trimer. These results suggest that the impact of mutations can be assessed, at least partially, by evaluating the interaction of the mutated peptides with the lipid membranes.

Dominant conformations of F19W 3Aβ11–40 immersed in transmembrane DPPC lipid bilayer submerged in aqueous solution.  相似文献   

4.
The search for efficient inhibitors targeting Aβ oligomers and fibrils is an important issue in Alzheimer''s disease treatment. As a consequence, an accurate and computationally cheap approach to estimate the binding affinity for many ligands interacting with Aβ peptides is very important. Here, the calculated binding free energies of 30 ligands interacting with 12Aβ11–40 peptides using the linear interaction energy (LIE) approach are found to be in good correlation with experimental data (R = 0.79). The binding affinities of these complexes are also calculated by using free energy perturbation (FEP) and molecular mechanic/Poisson–Boltzmann surface area (MM/PBSA) methods. The time-consuming FEP method provides results with similar correlation (R = 0.72), whereas MM/PBSA calculations show very low correlation with experimental data (R = 0.27). In all complexes, van der Waals interactions contribute much more than electrostatic interactions. The LIE model, which is much less time-consuming than both the FEP and MM/PBSA methods, opens the door to accurate and rapid affinity prediction of ligands with Aβ peptides and the design of new ligands.

The efficient approach to estimate inhibitors targeting Aβ oligomers and fibrils is an important issue in Alzheimer''s disease treatment.  相似文献   

5.
We have shown that fullerene (C60) becomes soluble in water by mixing fullerene and amyloid β peptide (Aβ40) whose fibril structures are considered to be associated with Alzheimer''s disease. The water-solubility of fullerene arises from the generation of a nanosized complex between fullerene and the monomer species of Aβ40 (Aβ40-C60). The prepared Aβ40-C60 exhibits photo-induced activity with visible light to induce the inhibition of Aβ40 fibrillation and the cytotoxicity for cultured HeLa cells. The observed photo-induced phenomena result from the generation of singlet oxygen via photoexcitation, inducing oxidative damage to Aβ40 and HeLa cells. The oxidized Aβ40 following photoexcitation of Aβ40-C60 was confirmed by mass spectrometry.

We have shown that fullerene (C60) becomes soluble in water by mixing fullerene and amyloid β peptide (Aβ40) whose fibril structures are considered to be associated with Alzheimer''s disease.  相似文献   

6.
Raman spectroscopy assisted by localized plasmon resonances generating effective hot spots at the gaps between intertwined silver nanowires is herein adopted to unravel characteristic molecular motifs on the surface of Aβ42 misfolded oligomers that are critical in driving intermolecular interactions in neurodegeneration.

Unraveling characteristic structural determinants at the basis of Aβ42 oligomers'' neurotoxicity by a sub-molecular SERS investigation of their surface.  相似文献   

7.
Alzheimer''s disease (AD), a neurodegenerative disorder, is marked by the accumulation of amyloid-β (Aβ) and neuroinflammation which promote the development of AD. Geniposide, the main ingredient isolated from Chinese herbal medicine Gardenia jasminoides Ellis, has a variety of pharmacological functions such as anti-apoptosis and anti-inflammatory activity. Hence, we estimated the inflammatory cytotoxicity caused by Aβ25–35 and the neuroprotective effects of geniposide in HT22 cells. In this research, following incubation with Aβ25–35 (40 μM, 24 h) in HT22 cells, the methylthiazolyl tetrazolium (MTT) and lactate dehydrogenase (LDH) release assays showed that the cell survival rate was significantly decreased. In contrast, the reactive oxygen species (ROS) assay indicated that Aβ25–35 enhanced ROS accumulation and apoptosis showed in both hoechst 33342 staining and annexin V-FITC/PI double staining. And then, immunofluorescence test revealed that Aβ25–35 promoted p65 to transfer into the nucleus indicating p65 was activated by Aβ25–35. Moreover, western blot analysis proved that Aβ25–35 increased the expression of nitric oxide species (iNOS), tumor necrosis factor-α (TNF-α), cyclooxygenase-2 (COX-2) and interleukin-1β (IL-1β). Simultaneously, Aβ25–35 also promoted the expression of toll-like receptor 4 (TLR4), p-p65 and p-IκB-α accompanied with the increase in the level of beta-secretase 1 (BACE1) and caspase-3 which further supported Aβ25–35 induced apoptosis and inflammation. Fortunately, this up-regulation was reversed by geniposide. In conclusion, our data suggest that geniposide can alleviate Aβ25–35-induced inflammatory response to protect neurons, which is possibly involved with the inhibition of the TLR4/NF-κB pathway in HT22 cells. Geniposide may be the latent treatment for AD induced by neuroinflammation and apoptosis.

Alzheimer''s disease (AD), a neurodegenerative disorder, is marked by the accumulation of amyloid-β (Aβ) and neuroinflammation which promote the development of AD.  相似文献   

8.
Presenilin (PSEN) pathogenic mutations cause familial Alzheimer’s disease (AD [FAD]) in an autosomal-dominant manner. The extent to which the healthy and diseased alleles influence each other to cause neurodegeneration remains unclear. In this study, we assessed γ-secretase activity in brain samples from 15 nondemented subjects, 22 FAD patients harboring nine different mutations in PSEN1, and 11 sporadic AD (SAD) patients. FAD and control brain samples had similar overall γ-secretase activity levels, and therefore, loss of overall (endopeptidase) γ-secretase function cannot be an essential part of the pathogenic mechanism. In contrast, impaired carboxypeptidase-like activity (γ-secretase dysfunction) is a constant feature in all FAD brains. Significantly, we demonstrate that pharmacological activation of the carboxypeptidase-like γ-secretase activity with γ-secretase modulators alleviates the mutant PSEN pathogenic effects. Most SAD cases display normal endo- and carboxypeptidase-like γ-secretase activities. However and interestingly, a few SAD patient samples display γ-secretase dysfunction, suggesting that γ-secretase may play a role in some SAD cases. In conclusion, our study highlights qualitative shifts in amyloid-β (Aβ) profiles as the common denominator in FAD and supports a model in which the healthy allele contributes with normal Aβ products and the diseased allele generates longer aggregation-prone peptides that act as seeds inducing toxic amyloid conformations.Early-onset familial Alzheimer’s disease (AD [FAD]), starting before age 65, is mainly caused by mutations in the Presenilin 1/2 (PSEN1/2) or the amyloid precursor protein (APP) genes and represents less than 0.1% of the total AD cases (Campion et al., 1999). Although rare, FAD offers a unique model to gain insights into the molecular mechanisms and etiology of sporadic AD (SAD).PSEN is the catalytic subunit of the γ-secretase complex (De Strooper et al., 1998; Wolfe et al., 1999), an intramembrane multimeric protease involved in the processing of many type 1 transmembrane proteins; among them, the Notch receptors and APP have received much attention because of their association with crucial cell signaling events or with AD pathogenesis, respectively (for a review see Jurisch-Yaksi et al. [2013]). Nicastrin (Nct), PSEN enhancer 2 (Pen2), and anterior pharynx defective 1 (APH1) are, together with PSEN, essential components of the protease complex (De Strooper, 2003).More than 150 pathogenic mutations in PSEN1 have been reported so far (http://www.molgen.ua.ac.be/ADMutations); and notably, the vast majority are missense substitutions distributed throughout the primary structure of PSEN1. PSEN/γ-secretase hydrolyzes peptide bonds in a process called regulated intramembrane proteolysis, which allows translation of extracellular signals into the cell. Compelling evidence indicates that γ-secretase cuts membrane proteins sequentially: the first endopeptidase cleavage (ε) releases a soluble intracellular domain (ICD), which may translocate to the nucleus to regulate gene expression while the remaining N-terminal transmembrane domain (TMD) fragment is successively cut by the carboxypeptidase-like activity of γ-secretase (γ-cleavages). Endopeptidase products, either amyloid-β49 (Aβ49) or Aβ48, are then processed along two major product lines: Aβ49 → Aβ46 → Aβ43 → Aβ40 or Aβ48 → Aβ45 → Aβ42 → Aβ38 (Takami et al., 2009). Every γ-cleavage removes a short C-terminal peptide from the TMD, reducing its hydrophobicity and increasing the probability of release. Secretion of an N-terminal fragment into the extracellular/luminal space terminates this sequence (Qi-Takahara et al., 2005; Yagishita et al., 2008; Takami et al., 2009). Importantly, the efficiency of the endopeptidase cleavage determines ICD product levels, which acquires high physiological relevance in the case of the Notch substrate. The carboxypeptidase-like efficiency, the number of cuts per substrate, determines the length of the N-terminal products; the level of efficiency is pathologically very relevant in the case of the APP substrate, as lower efficiency results in the production of longer and more aggregation-prone Aβ peptides (Chávez-Gutiérrez et al., 2012).How mutations in the PSENs cause FAD remains a hotly debated topic in the field. Because FAD is an autosomal-dominant disorder (patients carry both healthy and mutant alleles), a major unknown in the discussion remains the role of the healthy allele and, to a lesser extent, the role of the brain environment on the total (normal + mutant proteases) γ-secretase activity. To what degree do normal and mutant complexes contribute to total γ-secretase activity in the patient brain? Does the healthy allele compensate for the disease allele effects? Despite their relevance, those questions have not been addressed.Only one group, i.e., Potter et al. (2013), have estimated the Aβ production kinetics in the FAD brain by measuring isotope-labeled Aβ peptides in the cerebrospinal fluid of patients (stable isotope-labeled kinetics [SILK]). Feeding the in vivo metabolic labeling patient data into a mathematical model, specifically generated for their approach, they described higher Aβ42 production rates in the central nervous system of PSEN mutation carriers (Potter et al., 2013). Accordingly, Potter et al. (2013) suggest that increments in Aβ42 play a decisive pathogenic role in AD.In contrast, a “revised” loss-of-function (LOF) hypothesis has recently been proposed by Xia et al. (2015). In this view, loss of PSEN/γ-secretase physiological cell signaling function causes neurodegeneration, whereas changes in Aβ peptides are only secondary byproducts that arise from but do not trigger the disease (Xia et al., 2015). The idea is only tenable if FAD-linked PSEN mutations exert a LOF effect on PSEN/γ-secretase and, in addition, a dominant-negative effect on the healthy PSEN allele (normal γ-secretase) in patients, a key part of this hypothesis (Heilig et al., 2013; Xia et al., 2015). However, it should be stressed that γ-secretase haploinsufficiency caused by nonsense, frameshift, and splice site mutations in genes coding for essential subunits of γ-secretase (Nct, Pen2, and PSEN) is pathogenic in nature; such haploinsufficiency causes a chronic inflammatory disease of hair follicles known as familial acne inversa. Most importantly, no clinical association between this disorder and AD has been reported (for a review see Pink et al. [2013]). Furthermore, if FAD-linked PSEN mutations were truly LOF mutations, resulting in “inactive” γ-secretase complexes, homozygous individuals for the disease allele would not be viable because of disturbances in Notch signaling during embryonic development. However, six individuals with homozygous PSEN1 E280A gene mutation have been identified (Kosik et al., 2015).An alternative view to both hypotheses is that pathogenic mutations in PSEN cause disease by qualitative shifts in Aβ profile production (γ-secretase dysfunction; Chávez-Gutiérrez et al., 2012). We have demonstrated that loss of endopeptidase activity is not necessarily observed in γ-secretase complexes containing PSEN1/2 FAD-linked mutations, but reduced carboxypeptidase-like efficiency (γ-secretase dysfunction) is the constant denominator. Furthermore, FAD PSEN mutations may affect the carboxypeptidase-like γ-secretase activity at multiple turnovers, resulting in increased Aβ43 and Aβ42 levels as well as in other longer Aβ peptides, such as Aβ45 and Aβ46 (Quintero-Monzon et al., 2011; Chávez-Gutiérrez et al., 2012; Fernandez et al., 2014). These data support a model in which relative, rather than absolute, changes in Aβ product profiles are at the basis of PSEN/γ-secretase–mediated pathogenicity. However, these findings were based on studies conducted in PSEN1/2-deficient MEFs, which does not fully recapitulate the in vivo heterozygous situation in the FAD patient’s brain.In the current study, we investigated processing of APP by the γ-secretase complex in postmortem human brain samples from FAD and SAD patients and healthy control subjects. Our investigation is the first to directly assess γ-secretase activity in brain material from FAD mutant carriers and to address how the FAD-linked mutant heterozygous situation in patients affects γ-secretase function in brain.  相似文献   

9.

OBJECTIVE

Ketosis-prone diabetes (KPD) is characterized by diabetic ketoacidosis (DKA) in patients lacking typical features of type 1 diabetes. A validated classification scheme for KPD includes two autoantibody-negative (“A−”) phenotypic forms: “A−β−” (lean, early onset, lacking β-cell functional reserve) and “A−β+” (obese, late onset, with substantial β-cell functional reserve after the index episode of DKA). Recent longitudinal analysis of a large KPD cohort revealed that the A−β+ phenotype includes two distinct subtypes distinguished by the index DKA episode having a defined precipitant (“provoked,” with progressive β-cell function loss over time) or no precipitant (“unprovoked,” with sustained β-cell functional reserve). These three A− KPD subtypes are characterized by absence of humoral islet autoimmune markers, but a role for cellular islet autoimmunity is unknown.

RESEARCH DESIGN AND METHODS

Islet-specific T-cell responses and the percentage of proinflammatory (CD14+CD16+) blood monocytes were measured in A−β− (n = 7), provoked A−β+ (n = 15), and unprovoked A−β+ (n = 13) KPD patients. Genotyping was performed for type 1 diabetes–associated HLA class II alleles.

RESULTS

Provoked A−β+ and A−β− KPD patients manifested stronger islet-specific T-cell responses (P < 0.03) and higher percentages of proinflammatory CD14+CD16+ monocytes (P < 0.01) than unprovoked A−β+ KPD patients. A significant relationship between type 1 diabetes HLA class II protective alleles and negative T-cell responses was observed.

CONCLUSIONS

Provoked A−β+ KPD and A−β− KPD are associated with a high frequency of cellular islet autoimmunity and proinflammatory monocyte populations. In contrast, unprovoked A−β+ KPD lacks both humoral and cellular islet autoimmunity.Ketosis-prone diabetes (KPD), characterized by presentation with diabetic ketoacidosis (DKA) in patients lacking the typical features of autoimmune type 1 diabetes, is a heterogeneous syndrome (1,2). A validated classification scheme for KPD, based on the presence or absence of β-cell autoantibodies (“A+” or “A−”) and presence or absence of β-cell functional reserve (“β+” or “β−”) (3) includes two autoantibody-negative A− phenotypic forms: “A−β−” (lean, early onset, lacking β-cell functional reserve) and “A−β+” (obese, late onset, with substantial β-cell functional reserve after the index episode of DKA). Long-term longitudinal follow-up of a large cohort of A−β+ KPD patients has revealed that this phenotype comprises two distinct subtypes distinguished by whether the index DKA episode had a defined precipitant (“provoked” A−β+) or no precipitant (“unprovoked” A−β+) (4). Provoked A−β+ KPD patients have progressive loss of β-cell function after initial recovery from the DKA episode, with relapse to insulin dependence, no sex predominance, and an increased frequency of the HLA class II alleles DQB1*0302 and DRB1*04 associated with susceptibility to autoimmune type 1 diabetes. In contrast, unprovoked A−β+ KPD patients have sustained preservation of β-cell function after recovery from the DKA episode, prolonged insulin independence, male predominance, and an increased frequency of the protective allele DQB1*0602 (4). The unique clinical features and natural histories of these two subtypes of A−β+ KPD patients suggest distinctive underlying pathophysiologic processes for each. Although an underlying “occult” autoimmune element is suggested in the provoked A−β+ subtype by progressive β-cell loss and the presence of type 1 diabetes–associated HLA susceptibility alleles, the unprovoked A−β+ subtype could represent a truly nonautoimmune syndrome of KPD.We have previously shown that the T cells of a significant proportion of individuals with an apparent phenotype of type 2 diabetes react strongly to islet antigens, despite lacking β-cell autoantibodies, and that this reactivity is associated with low C-peptide levels, indicating underlying cellular immune damage to β-cells (5,6). This finding expands the range of diabetic phenotypes—including those labeled as having “type 2” diabetes—with a potential pathophysiologic basis in islet autoimmunity. In the current study, we extended these findings to the unique, emerging forms of A− KPD. Specifically, we hypothesized that differences in cellular immune responses might distinguish the three A− KPD subtypes (A−β−, unprovoked A−β+, and provoked A−β+) with regard to a cellular autoimmune etiology. To test this hypothesis, we measured islet-specific T-cell responses using the validated cellular immunoblotting assay and islet autoantibody responses to determine the presence of islet autoimmunity in patients carefully phenotyped for these three KPD subtypes. We further assessed the percentage of proinflammatory (CD14+CD16+) monocytes in the peripheral blood of the three KPD subtypes.In healthy subjects, 90–95% of classical monocytes express high levels of the cell surface marker CD14 (CD14+), without expression of CD16 (CD16−). Inflammation and infection are associated with the emergence of a distinct monocyte population characterized by coexpression of CD14 and CD16 (CD14+CD16+). CD14+CD16+ monocytes secrete high levels of proinflammatory tumor necrosis factor-α and low levels of anti-inflammatory interleukin-10, leading to their designation as proinflammatory monocytes (7).Our results demonstrate that 1) provoked A−β+ KPD, associated with progressive loss of β-cell function, is associated with a high frequency of cellular islet autoimmunity and proinflammatory monocyte populations; 2) unprovoked A− KPD is a distinct syndrome of reversible β-cell dysfunction lacking evidence of humoral or cellular islet autoimmunity; and 3) a substantial proportion of A−β− KPD patients, who resemble patients with type 1 diabetes but lack autoantibodies, have evidence of cellular islet autoimmunity.  相似文献   

10.
The apolipoprotein E (APOE) ε4 allele is the strongest genetic risk factor for Alzheimer’s disease (AD). The influence of apoE on amyloid β (Aβ) accumulation may be the major mechanism by which apoE affects AD. ApoE interacts with Aβ and facilitates Aβ fibrillogenesis in vitro. In addition, apoE is one of the protein components in plaques. We hypothesized that certain anti-apoE antibodies, similar to certain anti-Aβ antibodies, may have antiamyloidogenic effects by binding to apoE in the plaques and activating microglia-mediated amyloid clearance. To test this hypothesis, we developed several monoclonal anti-apoE antibodies. Among them, we administered HJ6.3 antibody intraperitoneally to 4-mo-old male APPswe/PS1ΔE9 mice weekly for 14 wk. HJ6.3 dramatically decreased amyloid deposition by 60–80% and significantly reduced insoluble Aβ40 and Aβ42 levels. Short-term treatment with HJ6.3 resulted in strong changes in microglial responses around Aβ plaques. Collectively, these results suggest that anti-apoE immunization may represent a novel AD therapeutic strategy and that other proteins involved in Aβ binding and aggregation might also be a target for immunotherapy. Our data also have important broader implications for other amyloidosis. Immunotherapy to proteins tightly associated with misfolded proteins might open up a new treatment option for many protein misfolding diseases.Accumulation of amyloid β (Aβ) peptide in the brain is hypothesized to initiate pathogenic cascades that lead to Alzheimer’s disease (AD; Golde et al., 2011; Holtzman et al., 2011). Therapeutic strategies to inhibit Aβ accumulation, such as Aβ immunotherapy, are currently being tested as disease-modifying therapies (Golde et al., 2011; Holtzman et al., 2011). Sequential proteolytic cleavages of Aβ precursor protein (APP) by β- and γ-secretase produce the Aβ peptide. Although rare, early-onset AD mutations in the APP, presenilin 1 (PS1), and PS2 genes have supported the critical role of Aβ in AD pathogenesis, the ε4 allele of apolipoprotein E (APOE) gene is the strongest genetic risk factor for the common late-onset form of AD, and the ε2 allele is protective (Kim et al., 2009a). Among several mechanisms proposed to explain the effects of apoE on AD pathogenesis, prevailing evidence suggests that apoE’s effect on Aβ accumulation may be the major mechanism (Kim et al., 2009a). ApoE appears to affect Aβ accumulation by modulating both Aβ aggregation and clearance (Ma et al., 1994; Wisniewski et al., 1994; Castano et al., 1995; Sadowski et al., 2006; Deane et al., 2008; Jiang et al., 2008; Castellano et al., 2011). Along with other amyloid-associated proteins, apoE is found in amyloid plaques (Namba et al., 1991; Wisniewski and Frangione, 1992). In this study, we hypothesized that after peripheral administration of monoclonal anti-apoE antibodies, the small fraction of anti-apoE antibodies that enter the brain would bind to apoE in plaques and decrease Aβ accumulation by activating microglia-mediated clearance of Aβ aggregates, similar to what has been described for certain anti-Aβ antibodies (Bard et al., 2000; Brody and Holtzman, 2008).  相似文献   

11.
Alzheimer''s disease (AD) is a neurodegenerative malady associated with amyloid β-peptide (Aβ) aggregation in the brain. Metal ions play important roles in Aβ aggregation and neurotoxicity. Metal chelators are potential therapeutic agents for AD because they could sequester metal ions from the Aβ aggregates and reverse the aggregation. The blood–brain barrier (BBB) is a major obstacle for drug delivery to AD patients. Herein, a nanoscale silica–cyclen composite combining cyclen as the metal chelator and silica nanoparticles as a carrier was reported. Silica–cyclen was characterized by scanning electron microscopy (SEM), transmission electron microscopy (TEM), Fourier transform infrared (FT-IR) and dynamic light scattering (DLS). The inhibitory effect of the silica–cyclen nanochelator on Zn2+- or Cu2+-induced Aβ aggregation was investigated by using a BCA protein assay and TEM. Similar to cyclen, silica–cyclen can effectively inhibit the Aβ aggregation and reduce the generation of reactive oxygen species induced by the Cu–Aβ40 complex, thereby lessening the metal-induced Aβ toxicity against PC12 cells. In vivo studies indicate that the silica–cyclen nanochelator can cross the BBB, which may provide inspiration for the construction of novel Aβ inhibitors.

A BBB-passable nanoscale silica–cyclen chelator effectively reduces the metal-induced Aβ aggregates and related ROS, thereby decreasing the neurotoxicity of Aβ.  相似文献   

12.
Age-related aggregation of amyloid-β (Aβ) is an upstream pathological event in Alzheimer’s disease (AD) pathogenesis, and it disrupts the sleep–wake cycle. The amount of sleep declines with aging and to a greater extent in AD. Poor sleep quality and insufficient amounts of sleep have been noted in humans with preclinical evidence of AD. However, how the amount and quality of sleep affects Aβ aggregation is not yet well understood. Orexins (hypocretins) initiate and maintain wakefulness, and loss of orexin-producing neurons causes narcolepsy. We tried to determine whether orexin release or secondary changes in sleep via orexin modulation affect Aβ pathology. Amyloid precursor protein (APP)/Presenilin 1 (PS1) transgenic mice, in which the orexin gene is knocked out, showed a marked decrease in the amount of Aβ pathology in the brain with an increase in sleep time. Focal overexpression of orexin in the hippocampus in APP/PS1 mice did not alter the total amount of sleep/wakefulness and the amount of Aβ pathology. In contrast, sleep deprivation or increasing wakefulness by rescue of orexinergic neurons in APP/PS1 mice lacking orexin increased the amount of Aβ pathology in the brain. Collectively, modulation of orexin and its effects on sleep appear to modulate Aβ pathology in the brain.Age-related aggregation of amyloid-β (Aβ) is an upstream pathological event in Alzheimer’s disease (AD) pathogenesis (Holtzman et al., 2011; Sperling et al., 2011). As the accumulation and aggregation of Aβ in the brain is known to develop ∼10–15 yr before the initial symptoms of AD, understanding the factors that lead to Aβ aggregation are likely to be important in delaying the onset of this pathology and delaying/preventing AD (Holtzman et al., 2011; Sperling et al., 2011). Diverse lines of in vitro and in vivo studies have shown that synaptic activity and specifically synaptic vesicle release is coupled with presynaptic Aβ release (Kamenetz et al., 2003; Cirrito et al., 2008; Bero et al., 2011; Roh et al., 2012). In addition, sleep plays a role in the regulation of synaptic weight in the brain such that sleep appears to downscale the slow wave activity of the brain caused by accumulated load of synaptic potentiation during wakefulness (Vyazovskiy et al., 2009). Poor sleep quality and insufficient amounts of sleep have been noted in humans with preclinical evidence of AD (Roh et al., 2012; Ju and Holtzman, 2013; Ju et al., 2014). These observations suggest that understanding how integrated synaptic and network activity as measured by the sleep–wake cycle regulates Aβ may provide novel insights into AD pathogenesis. However, how the amount and quality of sleep affect Aβ aggregation is not yet well understood (Ju et al., 2014). Orexins (hypocretins) initiate and maintain wakefulness, and loss of orexin-producing neurons causes narcolepsy (de Lecea et al., 1998; Chemelli et al., 1999).Levels of soluble Aβ in the extracellular interstitial fluid (ISF) of the hippocampus in mice are dynamically and positively associated with minutes awake per hour and negatively associated with time asleep (Kang et al., 2009). In addition, the sleep–wake cycle also affects the Aβ pathology in the brain. A study in mice showed that intracerebral administration of orexin can acutely increase both wakefulness and Aβ levels and systemic treatment with an orexin receptor antagonist decreased Aβ deposition in amyloid precursor protein (APP) transgenic mouse models (Kang et al., 2009). Pharmacological experiments suggest that orexin and the sleep–wake cycle appear to be related to regulation of Aβ levels. We sought to determine for the first time whether genetic manipulation of orexin has similar effects as pharmacological manipulation and, importantly, whether orexin is influencing Aβ levels and Aβ pathology directly via orexin signaling or indirectly via its effects on the sleep–wake cycle.  相似文献   

13.
Alzheimer''s disease (AD) is the most common form of neurodegenerative disease currently. It is widely accepted that AD is characterized by the self-assembly of amyloid beta (Aβ) peptides. The human glutaminyl cyclase (hQC) enzyme is characterized by association with Aβ peptide generation. The development of hQC inhibitors could prevent the self-aggregation of Aβ peptides, resulting in impeding AD. Utilizing structural knowledge of the hQC substrates and known hQC inhibitors, new heterocyclic and peptidomimetic derivatives were synthesized and were able to inhibit the hQC enzyme. The inhibiting abilities of these compounds were evaluated using a fluorometric assay. The binding mechanism at the atomic level was estimated using molecular docking, free energy perturbation, and quantum chemical calculation methods. The predicted log(BBB) and human intestinal absorption values indicated that these compounds are able to permeate the blood–brain barrier and be well-absorbed through the gastrointestinal tract. Overall, 5,6-dimethoxy-N-(3-(5-methyl-1H-imidazol-1-yl)propyl)-1H-benzo[d]imidazol-2-amine (1_2) was indicated as a potential drug for AD treatment.

Rational design of new hQC inhibitors.  相似文献   

14.
The onset of Alzheimer''s disease (AD) is associated with the presence of neurofibrillary pathology such as amyloid β (Aβ) plaques. Different therapeutic strategies have focused on the inhibition of Aβ aggregate formation; these pathological structures lead to neuronal disorder and cognitive impairment. Fullerene C60 has demonstrated the ability to interact and prevent Aβ fibril development; however, its low solubility and toxicity to cells remain significant problems. In this study, we synthesized, characterized and compared diethyl fullerenemalonates and the corresponding sodium salts, adducts of C60 bearing 1 to 3 diethyl malonyl and disodium malonyl substituents to evaluate the potential inhibitory effect on the aggregation of Aβ42 and their biocompatibility. The dose-dependent inhibitory effect of fullerenes on Aβ42 aggregation was studied using a thioflavin T fluorescent assay, and the IC50 value demonstrated a low range of fullerene concentration for inhibition, as confirmed by electron microscopy. The exposure of neuroblastoma to fullerenemalonates showed low toxicity, primarily in the presence of the sodium salt-adducts. An isomeric mixture of bisadducts, trisadducts and a C3-symetrical trisadduct demonstrated the highest efficacy among the tests. In silico calculations were performed to complement the experimental data, obtaining a deeper understanding of the Aβ inhibitory mechanism; indicating that C3-symetrical trisadduct interacts mainly with 1D to 16K residues of Aβ42 peptide. These data suggest that fullerenemalonates require specific substituents designed as sodium salt molecules to inhibit Aβ fibrillization and perform with low toxicity. These are promising molecules for developing future therapies involving Aβ aggregates in diseases such as AD and other types of dementia.

Synthesis of new non toxic nanomaterials, with high anti-amyloid fibrils formation effect, in vitro and in silico.  相似文献   

15.
Alzheimer''s disease (AD) is an extremely complex disease, characterized by several pathological features including oxidative stress and amyloid-β (Aβ) aggregation. Blockage of Aβ-induced injury has emerged as a potential therapeutic approach for AD. Our previous efforts resulted in the discovery of Monascus pigment rubropunctatin derivative FZU-H with potential neuroprotective effects. This novel lead compound significantly diminishes toxicity induced by Aβ(1-42) in Neuro-2A cells. Our further mechanism investigation revealed that FZU-H inhibited Aβ(1-42)-induced caspase-3 protein activation and the loss of mitochondrial membrane potential. In addition, treatment of FZU-H was proven to attenuate Aβ(1-42)-induced cell redox imbalance and Tau hyperphosphorylation which caused by okadaic acid in Neuro-2A cells. These results indicated that FZU-H shows promising neuroprotective effects for AD.

Monascus pigment rubropunctatin derivative FZU-H shows promising neuroprotective effects for AD.  相似文献   

16.
17.
There is extensive evidence that cholesterol and membrane lipids play a key role in Alzheimer disease (AD) pathogenesis. Cyclodextrins (CD) are cyclic oligosaccharide compounds widely used to bind cholesterol. Because CD exerts significant beneficial effects in Niemann-Pick type C disease, which shares neuropathological features with AD, we examined the effects of hydroxypropyl-β-CD (HP-β-CD) in cell and mouse models of AD. Cell membrane cholesterol accumulation was detected in N2a cells overexpressing Swedish mutant APP (SwN2a), and the level of membrane cholesterol was reduced by HP-β-CD treatment. HP-β-CD dramatically lowered the levels of Aβ42 in SwN2a cells, and the effects were persistent for 24 h after withdrawal. 4 mo of subcutaneous HP-β-CD administration significantly improved spatial learning and memory deficits in Tg19959 mice, diminished Aβ plaque deposition, and reduced tau immunoreactive dystrophic neurites. HP-β-CD lowered levels of Aβ42 in part by reducing β cleavage of the APP protein, and it also up-regulated the expression of genes involved in cholesterol transport and Aβ clearance. This is the first study to show neuroprotective effects of HP-β-CD in a transgenic mouse model of AD, both by reducing Aβ production and enhancing clearance mechanisms, which suggests a novel therapeutic strategy for AD.There is extensive evidence implicating cholesterol in the pathogenesis of AD. Epidemiological studies have shown that hypercholesterolemia is an important risk factor for AD (Kivipelto et al., 2001; Puglielli et al., 2003). Furthermore, higher cholesterol levels are observed in the brains of AD patients (Xiong et al., 2008). Cholesterol is a major component of the lipid bilayer membrane, where β- and γ-secretases are located. β-C-terminal fragments (β-CTFs), which are the products of β cleavage of the amyloid precursor protein (APP), directly bind cholesterol (Barrett et al., 2012). Increased cellular cholesterol increases the association of APP with β- and γ-secretases in cholesterol-rich lipid rafts, which stimulates cleavage of APP into Aβ (Lee et al., 1998; Ehehalt et al., 2003; Xiong et al., 2008; Kodam et al., 2010; Vetrivel and Thinakaran, 2010).Cholesterol homeostasis is tightly regulated by feedback mechanisms: sterol regulatory element binding proteins (SREBPs) and liver X receptors (LXRs) regulate the expression of genes that control the uptake, synthesis, and export of cholesterol. SREBPs regulate the expression of genes encoding HMG-CoA reductase and HMG-CoA synthase, which play essential roles in cholesterol synthesis (Brown and Goldstein, 1997; Bełtowski, 2008; Sato, 2010). LXRs activate genes encoding the ATP-binding cassette subfamily A/G transporters (ABCA/Gs), which control cholesterol efflux across the plasma membrane to apolipoprotein acceptors, initiating the reverse cholesterol transport pathway (Schmitz and Drobnik, 2002; Jessup et al., 2006). Studies in mouse models of AD indicate that ABCA1 plays an important role in the induction of ApoE lipidation, in facilitating Aβ transport to ApoE-containing carriers for clearance, and in suppression of Aβ generation. Disruption of ABCA1 leads to increased insoluble Aβ and parenchymal amyloid plaques, with no change in APP processing (Hirsch-Reinshagen et al., 2005; Koldamova et al., 2005a; Wahrle et al., 2005; Kim et al., 2007; Cramer et al., 2012). Activation of LXRs produces behavioral improvement and removal of fibrillar Aβ in transgenic mouse models expressing mutant APP (Koldamova et al., 2005b; Terwel et al., 2011).Cyclodextrins (CDs) are a family of cyclic polysaccharide compounds that can extract cholesterol from cultured cells (Zidovetzki and Levitan, 2007; Coskun and Simons, 2010; Di Paolo and Kim, 2011). In transgenic mouse models of the lipid storage disease Niemann-Pick type C1 (NPC1), subcutaneous administration of CDs improved cholesterol trafficking, corrected abnormal cholesterol metabolism, and prevented neurodegenerative changes in NPC1−/− mice (Liu et al., 2008, 2010; Zhang et al., 2009; Ramirez et al., 2010; Aqul et al., 2011). Notably, intriguing parallels exist between NPC1 and AD, including neurofibrillary tangles and prominent lysosomal system dysfunction (Nixon, 2004; Liu et al., 2010). A recent study showed that CDs reversed attenuated behavioral and pathological abnormalities in a mouse model generated by crossing NPC1+/− mice with mutant APP transgenic mice (Maulik et al., 2012). It is therefore plausible that CDs may also benefit AD (Aqul et al., 2011). In vitro studies showed that CDs reduce membrane cholesterol levels and reduce both BACE and γ-secretase activities (Hartmann et al., 2007), leading to additive effects in reducing amyloid-β production (Simons et al., 1998).In this study, we examined the hydroxypropyl form of β-CD (HP-β-CD) in Tg19959 mice, which overexpress human APP with the Swedish and Indiana familial AD mutations. These mice robustly develop Aβ pathology and memory deficits at 3–4 mo of age (Yao et al., 2010). We injected the mice subcutaneously with HP-β-CD starting at 7 d of age, twice weekly until 4 mo of age. Treatment with HP-β-CD significantly reduced Aβ burden and phosphorylated tau pathology, and improved memory in the Morris water maze. HP-β-CD activated expression of genes, including ABCA1, suggesting that HP-β-CD increases cholesterol transport and enhances Aβ clearance.  相似文献   

18.
Fourteen previously undescribed naturally occurring C-23 carboxylated triterpenoids, stewartiacids A–N (1–14), were isolated and characterized from the twigs and leaves of the ornamental and medicinal plant Stewartia sinensis (Chinese Stewartia), a ‘vulnerable’ species endemic to China. The new structures were elucidated on the basis of spectroscopic data, single crystal X-ray diffraction, and electronic circular dichroism (ECD) analyses. Stewartiacids A (1) and B (2) are isoursenol derivatives. Stewartiacid C (3) is a 12-oxo-γ-amyrin analogue. Both isoursenol and γ-amyrin derivatives are quite rare in nature. Stewartiacids D (4) and E (5) are 13,27-cycloursane-type compounds. Stewartiacids K (11) and L (12) are ursane-type triterpene and phenylpropanol adducts built through a 1,4-dioxane ring, which are also seldom reported in the literature. The rest are common C-23 carboxylated ursane-type (6–10) and oleanane-type (13, 14) pentacyclic triterpenoids. Stewartiacids G (7), K (11), and L (12) showed moderate inhibitory effects against ATP-citrate lyase (ACL), with IC50 values of 12.5, 2.8, and 10.6 μM, respectively. Stewartiacid K (11) also exhibited moderate inhibition (IC50: 16.8 μM) of NF-κB.

Fourteen new C-23 carboxylated triterpenoids (stewartiacids A–N, 1–14) were obtained from the ‘vulnerable’ plant Stewartia sinensis. 11 and 12 showed inhibitory effects against ACL.  相似文献   

19.
A curcumin derivative conjugated with Gd-DO3A (Gd-DO3A-Comp.B) was synthesised as an MRI contrast agent for detecting the amyloid-β (Aβ) fibrillation process. Gd-DO3A-Comp.B inhibited Aβ aggregation significantly and detected the fibril growth at 20 μM of Aβ with 10 μM of probe concentration by T1-weighted MR imaging.

A curcumin derivative conjugated with Gd-DO3A (Gd-DO3A-Comp.B) was developed to significantly inhibit the amyloid-β (Aβ) aggregation and detect the fibril growth by T1-weighted MR imaging.

A significant increase of Alzheimer''s disease (AD) patients urges the development of therapeutic and diagnostic technology.1 As with the therapeutic development, diagnostic technology also faces several obstacles. To date, the definite diagnosis of AD relies on the histopathological data of post-mortem.2,3 The non-invasive imaging technology targeting AD biomarkers such as amyloid β (Aβ) could provide phenotypical diagnostics, although the development of Aβ probes still remains challenging. Several contrast agents for single photon emission computed tomography (SPECT) and positron emission tomography (PET) such as Florbetapir-18F and Pittsburgh compound-B ([11C]PiB) were developed as efficient tracers in mild cognitive impairment patients.4,5 However, PET- and SPECT-based diagnostics require injection of radioactive probes, which cannot be measured frequently due to radiation exposure and limited availability of facilities. They also provide limited information on the anatomic profile of biomarkers due to their low spatial resolution and imprecise microscopic localization.6 In contrast, magnetic resonance imaging (MRI) contrast agents could quantify the Aβ accumulation in the anatomic brain image.7Several reported MRI contrast agents using gadolinium (Gd) complexes demonstrate potential use of Aβ detection. A clinically approved contrast agent, Gd(iii) diethylenetriaminepentaacetic acid (Gd-DTPA) complex accumulates in brain after opening the blood–brain barrier (BBB) by using mannitol and detects Aβ deposits in the mice AD-model.8 To improve the selectivity, Gd complexes were conjugated with compounds binding to Aβ such as Pittsburgh compound B (Gd-DO3A-PiB) which also serves as an approach for increasing MRI sensitivity.9,10 An α,β-unsaturated ketone compound curcumin has been widely reported as an Aβ probe due to its ability to bind the hydrophobic site of Aβ.11,12 Allen et al. firstly reported the direct conjugation of curcumin with Gd-DTPA which binds to Aβ with four times higher relaxivity than free Gd-DTPA.13 Furthermore, a polymalic acid-based nanoparticle covalently linked with curcumin and Gd-DOTA could also detect Aβ in human brain specimen by MRI.14 These previous studies demonstrate that the curcumin structure has significant potential for the development of MRI contrast agents for AD diagnosis.Previously, we reported a curcumin derivative, compound B, possesses 100-times stronger inhibitory activity of Aβ aggregation than curcumin on the basis of thioflavin T (ThT) competitive binding assay.15,16 According to this result, we designed curcumin-based Gd probes for the detection and inhibition of Aβ (Fig. 1A–C). We hypothesized that these probes could accelerate proton longitudinal relaxation depending on the fibrillation stage of Aβ, because molecular tumbling rate of the Gd complexes becomes slower (Fig. 1A).17 As a result, the probes permit the detection of Aβ by longitudinal relaxation time (T1)-weighted imaging. This mechanism could also be utilized to estimate the inhibitory activity of the probes by T1-based analysis (Fig. 1B). The curcumin and compound B were directly conjugated with the macrocyclic DO3A ligand through the propylamine linker to obtain Gd-DO3A-Cur and Gd-DO3A-Comp.B, respectively (Fig. 1C).Open in a separate windowFig. 1(A) A probe concept that produces T1 in a dependent manner of Aβ fibrillation process. (B) Inhibitor-based probes that cause moderate T1 decreases due to inhibitory activity of fibrillation. (C) The chemical structures of the synthesized Gd probes for Aβ detection and inhibition.Gd-DO3A-Cur and Gd-DO3A-Comp.B were synthesized according to Scheme 1 (detail in Scheme S1, ESI). The compound 5a and 5b, which have asymmetric curcumin derivatives containing carboxylic acid group, were synthesized by three step reactions. Amide bond formation with DO3A(tBu)3-propylamine ligand18 by condensation reaction afforded compound 7a and 7b. The tert-butyl groups were deprotected by trifluoroacetic acid producing compound 8a and 8b. The complexation was performed with GdCl3·6H2O by adjusting the reaction pH to 7, giving 43 and 41% yields of Gd-DO3A-Cur and Gd-DO3A-Comp.B, respectively. The T1 relaxivities (r1) of the curcumin-based Gd probes were estimated by T1 measurement using a 1 tesla NMR relaxometry (Fig. S1, ESI). For the comparison, we synthesized Gd-DO3A-Chal which is a reported probe for Aβ.19 The r1 of Gd-DO3A-Comp.B, Gd-DO3A-Cur, and Gd-DO3A-Chal were 7.1, 6.1 and 5.3 mM−1 s−1, respectively. These r1 values are higher than that of clinically approved Gd-DOTA (3.9 mM−1 s−1).20 The molecular weight of Gd-DO3A-Comp.B and Gd-DO3A-Cur is almost two times larger than that of Gd-DOTA. Because the r1 increases approximately linearly with molecular weight in low magnetic field,17 the high r1 values of Gd-DO3A-Comp.B and Gd-DO3A-Cur might be mainly attributed to their high rotational correlation time, rather than the high number of coordinated water molecules. The r1 of Gd-DO3A-Chal was comparable to the value reported previously.19Open in a separate windowScheme 1Synthetic scheme of Gd-DO3A-Cur and Gd-DO3A-Comp.B. (a) B(OH)3, morpholine, DMF, 100 °C, 10 min. (b) 3a/3b, B(OH)3, morpholine, DMF, 100 °C, 10 min. (c) TFA, DCM. (d) DO3A(tBu)3-propylamine ligand, PyBOP, HOBt, Et3N, DMF. (e) 7a/7b, TFA, DCM. (f) GdCl3·6H2O, NaOH, H2O.We evaluated the inhibitory effect of three probes toward Aβ aggregation by Congo red assay.21 After 24 h incubation of 20 μM Aβ with 10 μM probe, Gd-DO3A-Comp.B showed the lowest fluorescence intensity, indicating the strongest inhibitory activity followed by Gd-DO3A-Cur (Fig. 2A). As the comparison, the reported MRI agents, Gd-DO3A-Chal showed slight inhibitory activity. The inhibitory effect was further evaluated by transmission electron microscopy (TEM) with negative staining (Fig. 2B). In the absence of the probes, Aβ formed huge and massive fibril similar to the typical morphology of Aβ fibril.22 The TEM images of Aβ with Gd-DO3A-Comp.B showed the presence of white spheres below 10 nm, demonstrating that Gd-DO3A-Comp.B strongly inhibits Aβ aggregation. In fact, the fibril growth stopped at a stage of oligomer formation. Lower inhibitory activity of Gd-DO3A-Cur was also found to provide a shortened worm-like fibril, which is the typical morphology of Aβ exposed to curcumin.23 In contrast, the small amount of white spheres and partial fibril disruption were found in the image of Aβ with Gd-DO3A-Chal. In comparison with a reported Gd-DTPA-curcumin possessing inhibitory activity starting at 50 μM, Gd-DO3A-Comp.B possessed stronger inhibition of Aβ aggregation at 10 μM.24 The MTT assay using Neuro 2a cells showed that IC50 of Gd-DO3A-Cur and Gd-DO3A-Comp.B. were more than 500 μM, indicating that these compounds did not possess significant cytotoxicity (Fig. S2, ESI).Open in a separate windowFig. 2Inhibitory effect of the Gd probes toward Aβ aggregation measured by Congo red assay (A) and negative staining TEM images (B). The Gd probes were co-incubated with monomeric Aβ for 24 h in PBS at pH 7.4. [Gd] = 10 μM, [Aβ] = 20 μM. Scale bars = 100 nm.To detect fibrillation process by NMR relaxometry, we measured T1 of the probe mixture with Aβ which were pre-incubated for 1, 3, 6, 12, and 24 h to make it form the fibrils of different growth stages (Fig. 3A and B). The T1 of Gd-DO3A-Comp.B solution decreased with pre-incubation time of Aβ, demonstrating that the Gd-DO3A-Comp.B can detect Aβ fibril depending on the growth stage (Fig. 3B). Lower T1 involved with Aβ growth could be caused by the reduction in tumbling rate of the Gd complex.25 We also co-incubated the probes with the Aβ monomer and monitored T1 changes over the incubation time (Fig. 3A, B and S3, ESI). Interestingly, the Gd-DO3A-Comp.B did not cause significant T1 decreases even after 24 hours co-incubation with Aβ monomers, demonstrating that Gd-DO3A-Comp.B has a strong inhibitory effect on fibril formation and the inhibition can be monitored by T1 measurement (Fig. 3B). The inhibitory effect was consistent with the results of Congo red assay and TEM (Fig. 2). On the other hand, the time-dependent increases of T1 were observed in Gd-DO3A-Chal and Gd-DO3A-Cur. This might be because these two probes were buried in the hydrophobic pocket as Aβ fibril grew up and fewer water molecules permitted access to the Gd ions. It is also possible that these probes have lower binding affinity, especially for matured fibril, and require higher concentrations to produce significant T1 changes.26 These probe did not produce the significant ΔT1 between monomer and fibril samples (Fig. 3B and S3, ESI), although they showed little inhibition in Congo red assay and TEM (Fig. 2).Open in a separate windowFig. 3(A) Experimental design of T1-based detection of Aβ fibrillation and inhibition by using the Gd probes. (B) T1 changes of the Gd probe solutions with pre-incubated fibrils and monomers in PBS at pH 7.4 (mean ± SEM, n = 3). [Gd] = 10 μM, [Aβ] = 20 μM.The feasibility of the Gd probes was further evaluated by in vitro MRI measurement using a 1 tesla scanner. The T1-weighted images showed that Gd-DO3A-Comp.B produced slight T1 signal increases with Aβ monomers for 2 and 24 h (Fig. 4A and B). More significant signal increases were observed in the Gd-DO3A-Comp.B with Aβ fibril pre-incubated for 24 h (Fig. 4C). In contrast, Gd-DO3A-Chal and Gd-DO3A-Cur did not show significant signal changes in the presence of Aβ monomers or fibrils (Fig. 4A–C). These results were mostly consistent with the T1 profile measured by NMR (Fig. 3). Compared to the previously reported Gd-DO3A-Chal that required 100 μM of the probe concentration to detect the equimolar Aβ,19 Gd-DO3A-Comp.B could detect five-times lower concentration of Aβ (20 μM) with ten-times lower probe concentration (10 μM). Therefore, Gd-DO3A-Comp.B could be promising to further develop highly sensitive diagnostic MRI contrast agents of AD.Open in a separate windowFig. 4 T 1-weighted images of the Gd probe solutions in the presence of monomeric Aβ at 2 h incubation (A), monomeric Aβ at 24 h incubation (B), and Aβ fibrils pre-incubated for 24 h (C). Incubation was conducted in PBS at pH 7.4.In conclusion, we synthesized the curcumin-based Gd probes which enabled the detection and inhibition of Aβ fibril formation. Gd-DO3A-Comp.B allowed for the highly sensitive detection of Aβ fibril by the T1 measurement. Moreover, the inhibitory activity could be estimated by T1 measurement, because Gd-DO3A-Comp.B decreased T1 depending on the growth stage of Aβ fibril formation. Such unique modality would be useful not only for the diagnostics but also for the direct evaluation of the therapeutic efficacy in vivo. For the future application, it would be important to combine with BBB penetration methods targeting the brain such as transient opening of the BBB using focused ultrasound or mannitol injection.27,28  相似文献   

20.
The C-terminus fragment (Val-Val-Ile-Ala) of amyloid-β is reported to inhibit the aggregation of the parent peptide. In an attempt to investigate the effect of sequential amino-acid scan and C-terminus amidation on the biological profile of the lead sequence, a series of tetrapeptides were synthesized using MW-SPPS. Peptide D-Phe-Val-Ile-Ala-NH2 (12c) exhibited high protection against β-amyloid-mediated-neurotoxicity by inhibiting Aβ aggregation in the MTT cell viability and ThT-fluorescence assay. Circular dichroism studies illustrate the inability of Aβ42 to form β-sheet in the presence of 12c, further confirmed by the absence of Aβ42 fibrils in electron microscopy experiments. The peptide exhibits enhanced BBB permeation, no cytotoxicity along with prolonged proteolytic stability. In silico studies show that the peptide interacts with the key amino acids in Aβ, which potentiate its fibrillation, thereby arresting aggregation propensity. This structural class of designed scaffolds provides impetus towards the rational development of peptide-based-therapeutics for Alzheimer''s disease (AD).

Amidated C-terminal fragment, Aβ39–42 derived non-cytotoxic β-sheet breaker peptides exhibit excellent potency, enhanced bioavailability and improved proteolytic stability.

First reported by Alois Alzheimer in 1906, Alzheimer''s Disease (AD) is a progressive, neurodegenerative disorder with an irreversible decline in memory and cognition.1 It is commonly seen in elderly populations and is marked by two major histopathological hallmarks, amyloid-β (Aβ) plaques and neurofibrillary tangles (NFT).2 The number of patients suffering from AD has been increasing at an alarming rate. It is estimated that around 60 million patients will be suffering from AD by the end of 2020 and half of this patient population would require the care equivalent to that of a nursing home.3 Even after a century of its discovery, there has been no treatment that targets the pathophysiology of AD.4 The current treatment regimen includes acetylcholine-esterase inhibitors (AChEI) comprising of donepezil, rivastigmine and galantamine as well as N-methyl-d-aspartate (NMDA)-receptor antagonist, memantine. These provide only symptomatic relief to the patient. Most of the therapeutics that are currently being tested in clinical trials couldn''t proceed beyond phase II and phase III clinical trials due to lower efficacy in elderly patients, multiple side effects and limitations in their pharmacokinetic and pharmacodynamic profiles.5–9 This has created challenges on the societal and economic upfront.Literature analysis reveals that the soluble oligomeric Aβ species is the culprit for neurotoxicity. It interrupts normal physiological functioning of the human brain. The central hydrophobic fragment Aβ16–22 (KLVFFAE) and the C-terminus region fragment Aβ31–42 (IIGLMVGGVVIA) are responsible for controlling the aggregation kinetics of the monomeric species, wherein the later still remains relatively less explored.8,9 Our group is focused on development of peptidomimetic analogues for preventing Aβ aggregation. Studies on a complete peptide scan on the C-terminus region regions has already been published previously.10–12 In an attempt to enhance the biological efficacy of the previously designed scaffolds,12 we rationalized the use of sequential amino acid scan by modifying/replacing individual residues, as well as amide protection of the C-terminus on the lead tetrapeptide sequence (Val-Val-Ile-Ala) to enhance the proteolytic stability of the peptides.C-terminus amidated peptides were synthesized by microwave-assisted Fmoc-solid phase peptide synthesis protocol. Scheme 1 shows the general route for the synthesis of peptides employing Rink amide resin (detailed methodology is mentioned in the ESI, Section 1). These peptides were characterized using by analytical HPLC, 1H and 13C NMR, APCI/ESI-MS and HRMS.Open in a separate windowScheme 1Synthesis of tetrapeptide 12c using MW-assisted Fmoc-solid phase peptide synthesis protocol employing Rink amide resin. Reaction conditions: (i) 20% piperidine in DMF (7 mL), MW (40 W), 60 °C, 2 cycles – 1.5 & 3 min; (ii) 2, 4, 6, 8 (4 equiv.), TBTU (4 equiv.), HoBt (4 equiv.), DIPEA (5 equiv.), DMF (3.5 mL), MW (40 W), 60 °C, 13.5 min; (iii) TFA : TIPS : H2O (95 : 2.5 : 2.5), rt, 2.5 h; reaction monitoring was done by: UV measurement: Fmoc deprotection-dibenzofulvene adduct, Kaiser test: 1° amines; acetaldehyde test: 2° amines.Aggregation of Aβ42 results in accumulation of toxic species, which interact with neurons and hinder their functioning, leading to loss of memory and cognition. Inhibiting the process of Aβ42 aggregation would alleviate the neurotoxicity imparted in PC-12 cells; this was evaluated by MTT cell viability assay.13 Viability of untreated cells is considered 100%. Upon treatment with 2 μM of Aβ42, only 74% cells were found to be viable. Out of the total tested peptides, seven tetrapeptides 12a, 12c, 12f, 13e, 14b, 15b and 15f showed complete inhibition of Aβ42-induced toxicity by restoring the cell viability to 100%, at respective concentrations. Results for cell viability assay along with Thioflavin-T fluorescence assay for all the synthesized tetrapeptides has been summarized in
No.Test peptide sequenceaMTT cell viability assayThT-fluorescence assay
Test peptide concentration range (Aβ42: test peptide)
10 μM (1 : 5)4 μM (1 : 2)2 μM (1 : 1)10 μM (1 : 5)4 μM (1 : 2)2 μM (1 : 1)
% viable cellsb% inhibitiond
11Val-Val-Ile-Ala-OH (lead)93.690.578.966.260.633.9
11aVal-Val-Ile-Ala-NH282.289.386.452.152.758.8
12a d -Val-Val-Ile-Ala-NH281.6100.094.882.183.564.7
12b Phe-Val-Ile-Ala-NH292.195.796.277.666.251.9
12c d -Phe-Val-Ile-Ala-NH2100.095.298.4100.0100.0100.0
12d d -Pro-Val-Ile-Ala-NH292.083.582.839.852.960.8
12e Nva-Val-Ile-Ala-NH283.486.174.170.351.977.2
12f Aib-Val-Ile-Ala-NH261.395.1100.093.866.064.5
12g Gly-Val-Ile-Ala-NH275.494.392.198.164.584.4
13aVal-d-Val-Ile-Ala-NH287.675.274.715.746.589.3
13bVal-d-Ile-Ile-Ala-NH294.692.890.349.435.165.6
13cVal-Pro-Ile-Ala-NH283.196.093.018.516.331.8
13dVal-Aib-Ile-Ala-NH2100.093.075.145.350.052.1
13eVal-Phe-Ile-Ala-NH293.3100.079.357.261.7100.0
13fVal-d-Phe-Ile-Ala-NH299.792.994.262.566.275.8
14aVal-Val-d-Ile-Ala-NH269.676.579.721.825.349.6
14bVal-Val-Leu-Ala-NH286.3100.083.80.016.542.2
15aVal-Val-Ile-d-Ala-NH293.880.283.323.214.668.6
15bVal-Val-Ile-Aib-NH299.1100.071.935.142.024.9
15cVal-Val-Ile-Gly-NH290.588.580.75.431.862.7
15dVal-Val-Ile-Val-NH271.775.464.511.06.90.0
15eVal-Val-Ile-Leu-NH270.671.789.431.230.018.7
15fVal-Val-Ile-Ile-NH2100.097.078.648.20.00.0
16a Pro-Pro-Ile-Ala-NH277.760.774.614.218.324.7
4274.05
Controlc100.0
Open in a separate windowaAmino acid residue modified within the tetrapeptide sequence is indicated in bold.bCell viability studies were performed using MTT cell viability assay against PC-12 cells.cThe percentage of untreated cells was considered 100% (positive control); percentage cell viability was calculated for the cells incubated along with Aβ42 (2 μM) in absence (negative control) and presence of the test peptides in respective dose concentrations for 6 h. % of viable cells was calculated by the formula as 100 × [Aβ42 + test peptide OD570 − Aβ OD570/control OD570 − Aβ OD570]. In a subset of triplicate wells, standard deviation values ranged 1.81–4.72.dInhibition of Aβ42 aggregation was calculated by Thioflavin-T fluorescence assay. % relative fluorescence units (% RFU) exhibited by Aβ fibrils were considered as 100%. ThT dye incubated alone was considered as control and % RFU units were computed when Aβ42 was co-incubated with the test peptides for 24 h (λex 440 nm, λem 485 nm). % inhibition of ThT fluorescence was calculated by using the formula: 100 × [100 − (Aβ42 + test peptide RFU485 − control RFU485/Aβ42 RFU485 − control RFU485)]. In a subset of triplicate wells, SD values ranged 1.22–4.83. Data for both the experiments was recorded for triplicate samples and the readings were averaged (<5% variation).The quantitative evaluation of β-sheet structures within the amyloid fibrils is accessed by the fluorescence of the Thioflavin-T dye.14 Inhibiting amyloid fibrillation would reduce or eliminate such an enhancement in fluorescence. This concept is utilized to evaluate compounds that would prevent Aβ from aggregating.15–17 ThT fluorescence in the presence of Aβ42 alone was considered 100% and % relative fluorescence unit (% RFU) values were calculated for Aβ42 co-incubated with the respective inhibitor peptides. ThT incubated alone, exhibited % RFU of nearly 53.3% as compared to control solution without dye. Complete data of % inhibition of Aβ42 by the test peptides has been summarized in 17 Out of all the tested peptides, peptides 12c and 13e showed minimal enhancement in ThT fluorescence when co-incubated with equimolar concentrations of Aβ42. Peptides 12f and 12g showed >90% activity at five-fold excess dose concentrations. A comparative bar graph representation for the four most active peptides 12c, 12f, 12g and 13e has been depicted in Fig. 1A. To understand the % RFU values indicating the relative fluorescence of ThT and % inhibition exhibited by the most active test peptides 12c, 12f, 12g and 13e, values have been summarized in ESI, Tables S1 and S2. The observed RFU was close to that of the control wells where the dye incubated alone. Negligible increment of ThT fluorescence when the test peptides are co-incubated with Aβ42 peptide clearly indicates the inhibition of fibrillation. It also provides further support to the inhibition of Aβ42-induced neuronal toxicity as studied in the MTT assay. The % inhibition of Aβ42 aggregation as depicted by the ThT fluorescence assay is in accordance with the % cell viability data obtained by the MTT assay. Exact correlation cannot be established between the two because of the differential behavior of Aβ42 in the presence of a cellular environment as well as the treatment/incubation time for both the experiments.Open in a separate windowFig. 1Effect of most active test peptides on Aβ42 aggregation: bar plots depicting the decrease in % RFU of ThT dye when Aβ42 (2 μM) was co-incubated with test peptides at higher doses (A), and lower doses (B). Complete fluorescence was represented by the Aβ42 peptide incubated along with the dye (black) and dye control (grey) represents the dye incubated alone. Subsequent bars represent Aβ peptide co-incubated with the varying concentrations of inhibitor peptides for 24 h. Significance values indicated with respect to the Aβ peptides, *, p < 0.05; **, p < 0.01; ***, p < 0.001. (C) Dose dependent modulation of Aβ42 aggregation-induced-neurotoxicity in PC-12 cells exhibited by the test peptide 12c (black). (D) Concentration dependent % inhibition on Aβ42 aggregation mediated ThT fluorescence exhibited by the test peptide 12c (black). % inhibition of ThT fluorescence was calculated by using the formula: 100 × [100 − (Aβ42 + test peptide RFU485 − control RFU485/Aβ42 RFU485 − control RFU485)]. Readings (λex 440 nm, λem 485 nm) was recorded for triplicate samples from three individual experiments and the readings were averaged (<5% variation). Error bars represent mean ± SD (n = 3). Data were analyzed by one-way anova test.Peptides 12c and 12f that exhibited >98% cell viability at the lowest tested concentration of 2 μM were then evaluated at lower dose concentrations of 1.0 μM, 0.5 μM and 0.1 μM, against 2 μM of Aβ42 maintaining the ratios of 1 : 2, 1 : 4 and 1 : 20 (test peptide : Aβ42), respectively. The graphical plot of dose dependent modulation of Aβ42 aggregation-induced-neurotoxicity in PC-12 cells is depicted in Fig. 1C. Peptide 12c exhibited 78% inhibition of Aβ42 even at a lowest tested dose of 0.1 μM. A graphical representation of the % decrease in RFU has been depicted in Fig. 1B and dose dependent % inhibition of Aβ42 aggregation has been depicted in Fig. 1D. Excellent activities were exhibited when the test peptide is present in equimolar concentrations of Aβ42, indicating 1 : 1 inhibition of the parent peptide.Upon the basis of careful analysis of data reported herewith and results published earlier,10–12 a correlation between the amino acid residues within the tetrapeptide sequence and the exhibited activity was established. Replacement of the first residue, Val39 with hydrophobic residues (Phe and D-Phe) exhibited >90% inhibition. The replacement of Val40 with hydrophobic (Phe, D-Ile) or conformationally restricted amino acids (Pro, Aib) slightly enhanced the potency of the peptide. This holds true even in dual substitution at Val39 and Val40. Any replacement or modification yielded less active derivatives, indicating Ile41 is critical for activity. Small analogous amino acid is preferred at the Ala42 for retaining the activity. Amidation of the C-terminus results in enhancement of inhibition potential for some peptides. In some cases, a decrease in activity is also observed. A summary of the SAR is shown in Fig. 2. Understanding the sequential positioning of these residues would help us further develop derivatives with enhanced potency to inhibit Aβ42 aggregation.Open in a separate windowFig. 2Derived structure–activity-relationship of tetrapeptides.It is reported that there are two species of Aβ i.e.40 and Aβ42, present in the diseased brain.18–20 Evaluating the inhibitory activities of the test compounds on the aggregation of both the species would be of biological significance.21 Therefore, inhibitory potential of peptide 12c was evaluated on Aβ40. The relative increase in the ThT fluorescence when Aβ40 was incubated alone for 24 h was very less or similar to that of the control wells containing ThT (Fig. 3A). Further testing for 48 h and 72 h, respectively yielded no substantial results (ESI, Table S3).Open in a separate windowFig. 3Thioflavin-T fluorescence studies on Aβ species: % RFU exhibiting the effect of peptide 12c on aggregation of (A) Aβ40 (5 μM) and (B) Aβ40 & Aβ42 mixture (5 μM) mediated ThT fluorescence. Complete fluorescence was represented by Aβ40 and the 10 : 1 mixture of Aβ peptides incubated along with the dye (black) and dye incubated alone (grey). Subsequent bars represent the respective concentrations of inhibitor peptide 12c co-incubated with the corresponding Aβ peptides for 24 h. Readings (λex 440 nm, λem 485 nm) was recorded for triplicate samples from three individual experiments and the readings were averaged (<5% variation). Error bars represent mean ± SD (n = 3). Data were analyzed by one-way anova test. Significance values indicated with respect to the Aβ peptides, *, p < 0.05; **, p < 0.01; ***, p < 0.001.The minimal increase in fluorescence could be attributed to the slower nucleation rate and longer lag phase in the kinetics of Aβ40 fibril formation in comparison to Aβ42.22–24 Also, Aβ40 is relatively less neurotoxic and its aggregation propensity enhances in the presence of Aβ42, although the former is present in a ten-fold higher concentration. Thus, evaluation of the inhibitory activity of test peptide 12c on the mixtures of Aβ40 : Aβ42 in the ratio of 10 : 1 was performed. On incubation of a mixture of 5 μM of Aβ40 and 0.5 μM of Aβ42 (ratio of 10 : 1) the % relative increase in ThT fluorescence was comparatively higher than when Aβ40 was incubated alone (Fig. 3B). When compared to that of Aβ42 incubated alone as analyzed in the previous experiments, the fluorescence intensities were less. This gave us a clear indication that the aggregation propensity of Aβ40 is amplified in the presence of 0.1 equimolar Aβ42. On co-incubation of the test peptide 12c with the 5 μM mixture of Aβ40 : Aβ42 in the ratio of 10 : 1, substantial decrease in the fluorescence levels were observed (ESI, Table S4).As a prophylactic measurement, the potential ability of the test peptides to deform the aggregated Aβ42 was investigated.25,26 Monomeric Aβ42 was pre-incubated for a period of 24 h and fluorescence was measured. As anticipated, there was a marked increase in the fluorescence due the fibril state of Aβ42. Test peptide was added to each of the wells at the respective dose concentrations and readings were recorded at in a time dependent manner until 120 h of incubation. Time dependent % deformation of preformed fibrils in the presence of equimolar test peptide has been depicted in Fig. 4A. It could be observed that peptide 12c has the potential of deforming pre-aggregated Aβ42 fibrils (>35%) until 48 h of incubation. Also, peptide 12c significantly reduced the fluorescence of Aβ42 showing 57.5, 34.9 and 33.7% inhibition at 2, 1 and 0.5 μM, respectively after 24 h treatment. % inhibition and % RFU for time intervals of 24 h and 48 h has been provided in the ESI, Table S5.Open in a separate windowFig. 4Time dependent inhibition of Aβ42: (A) % deformation or disaggregation of preformed Aβ42 fibrils in presence of equimolar concentration of test peptide 12c as evaluated via ThT fluorescence assay. (B) Time and concentration dependent RFU comparison depicting the effect of individual tetrapeptide 12c on Aβ42 mediated-ThT fluorescence. % inhibition of ThT fluorescence was calculated by using the formula: 100 × [100 − (Aβ42 + test peptide RFU485 − control RFU485/Aβ42 RFU485 − control RFU485)]. Readings (λex 440 nm, λem 485 nm) was recorded for triplicate samples and the values were normalized to the ThT dye control and averaged (<5% variation). Data was interpreted from three individual experiments. Error bars represent mean ± SD (n = 3). Data were analyzed by one-way anova test.A time dependent ThT fluorescence assay was performed on the most active test peptide 12c at the similar concentrations of 2, 1 and 0.5 μM with Aβ42 (2 μM) for a period of 7 days. Readings were recorded at regular time intervals of 24 h each. Fig. 4B shows the decrease in the RFU values when Aβ42 was incubated in presence of peptide 12c. Aβ42 on incubation alone with the ThT dye showed an enhancement in the fluorescence of about 57% that could be attributed to the aggregation of the Aβ42 peptide. The fluorescence shown by the blank wells, wherein the dye incubated alone was considered as the control. Compared to the Aβ42 sample, very low values of fluorescence were observed in the presence of peptide 12c. % inhibition of Aβ42 aggregation exhibited by the test peptide has been summarized in ESI, Table S6. It can be summed that peptide 12c exhibits activity on preformed Aβ42 fibrils as well as inhibits Aβ42 aggregation until 120 h of treatment.ANS fluorescence assay was performed in complimentary to Thioflavin-T assay.27–29 The effect of test peptide inhibiting the process of Aβ42 aggregation as well as deformation of the preformed Aβ42 fibrils was evaluated. The relative fluorescence for Aβ42 fibrils was considered to be 100% and decrease in the % RFU when test peptides were co-incubated with Aβ42 was computed. The fluorescence emitted by binding of ANS to the test peptide itself was subtracted as the test peptide is hydrophobic. The results for relative decrease in fluorescence obtained in both the sub-experiments have been summarized in Fig. 5A. Upon co-incubation of Aβ42 and test peptide 12c in ratios 1 : 1 and 1 : 0.5, a marked decrease in the fluorescence intensity was observed, in comparison to that of the lowest tested concentration of 0.5 μM. These results are in accordance to the results seen in ThT fluorescence assay.Open in a separate windowFig. 5Additional fluorescence studies: (A) effect of varying concentration of tetrapeptide 12c on inhibition of Aβ42 fibril formation (Set 1) and on pre-aggregated fibrils of Aβ (Set 2). Complete fluorescence was represented by the Aβ42 (2 μM) incubated alone, monomeric (Set 1) and pre-aggregated t = 24 h (Set 2) and in the presence of respective concentrations of the test peptide 12c after 24 h. ANS dye incubated alone was considered as control and % RFU units for individual samples were computed by normalizing to the ANS dye control (λex 480 nm, λem 535 nm). Subsequent bars represent the % RFU of the respective concentrations of the inhibitor peptide 12c co-incubated with the differential states of Aβ42 peptide (2 μM) for 24 h. (B) Fluorescence spectrum showing effect of test peptide on Aβ42 aggregation and its interaction with GUVs. Fluorescence of Aβ42 alone at 0 h (green), 24 h (black); along with test peptides 12c (blue) after 24 h in the presence of GUVs (λex 480 nm, λem 400–600 nm). ANS dye incubated alone was considered as control and relative FL. Intensities for individual samples were computed by normalizing to the ANS dye control. (C) Intrinsic tyrosine fluorescence of Aβ42 during fibrillation and inhibition by test peptide 12c (λex 260 nm, λem 280–410 nm). Fluorescence of 5 μM Aβ42 (t = 0 h, green), 5 μM Aβ42 incubated alone (t = 24 h, black), Aβ42 co-incubated along with 5 μM of the test peptides, 12c (t = 24 h, blue). Readings was recorded for triplicate samples from three individual experiments and were averaged (<5% variation). Error bars represent mean ± SD (n = 3). Data were analyzed by one-way anova test.It is hypothesized that the aggregated soluble oligomeric form of Aβ42 interacts with the neuronal membranes by hampering cellular processes and exhibiting neurotoxicity.7 The design of the experiment was similar to the MTT test conditions, wherein giant unilamellar vesicles (GUVs) with composition mimicking the rat neuronal myelin were prepared (ESI, Section 5.1) and interaction of Aβ was evaluated by ANS fluorescence measurements.29–32 When Aβ42 was just added to the prepared GUVs (t = 0 h), ANS showed a good emission spectrum from 450 to 550 nm. After 24 h incubation of Aβ with the vesicles, no emission band was seen, indicating the absence of hydrophobic binding domain, thus no fluorescence.This could be attributed that the hydrophobic region entered the vesicles, providing no binding site for the dye. Emission intensities obtained for the vesicles alone was considered as blank and was subtracted from the readings obtained for both the time points. To evaluate the effects of incubation of test peptides and its inhibitory effect on Aβ42 aggregation, Aβ42 and test peptide 12c along with the GUVs was incubated for 24 h at 37 °C and the relative change in the fluorescence was observed. Readings for the test peptide incubated alone with the vesicles at the similar concentration were subtracted the final readings so as to obtain a comparable result for Aβ. On co-incubation of Aβ42 with 12c, two distinct observations were seen, primarily a visible emission spectrum and secondly a blue shift with a slight increase in the emission intensity (Fig. 5B). The emission spectrum indicates that the hydrophobic region was available for binding to ANS, thus indicating that Aβ was available in its monomeric form itself. The increase in emission intensity and the blue shift does suggest certain interactions between 12c and the full-length Aβ, which results into differential binding of ANS to the test peptide–Aβ complex (ESI, Fig. S2).Monitoring intrinsic Tyr fluorescence of Aβ42 during fibril formation and interaction with most active test peptides was also studied. Tyr has significantly lower quantum yield than Trp and is usually only used as an intrinsic fluorescent probe in Trp-lacking proteins or peptides, since energy transfer to Trp residues usually quenches the Tyr fluorescence. Tyrosine shows a typical emission band at 305 nm, which is red shifted to 340 nm due to resonance of the phenol to phenolate ion. We hypothesized that in aggregated state of Aβ42, stacking of phenolate ions of the tyrosine residues will produce slightly enhanced fluorescence in comparison to that of the monomeric or non-aggregated state. It could be clearly seen, in the overlay of fluorescence spectrum for Aβ42 incubated alone for 0 h (green) and 24 h (black), as well as in presence of 12c (blue, Fig. 5C) that the monomeric state of Aβ42 was retained in presence of the test peptides. A comparative bar plot analysis (ESI, Fig. S3) of the fluorescence response at 340 nm has been indicated to compare the change in observed fluorescence intensities.Since amyloid-β aggregation is preceded by the conformational transition towards increasing β-sheet structure, monitoring the content of β-sheet formation would therefore depict the effect of inhibitors on the aggregation of Aβ42.32,33 The effect of inhibitor peptides on the conformation of Aβ42 was assessed via CD spectroscopy.34–36 Spectra of equimolar concentrations of the individual test peptides were subtracted from the corresponding spectra of inhibitor peptides co-incubated Aβ42. Predicted values of conformation are summarized in ESI, Table S7. When incubated alone, Aβ42 exhibited a conformational transition from random coiling and turn to majorly β-sheet form. Initially, Aβ42 peptide majorly comprised of 49.4% β-sheet conformation. At the end of 24 h incubation period, β-sheet content increased to 66.4%, with the α-helix content increasing from 6.3% to 17.0%. In the presence of inhibitor peptide 12c, β-sheet form completely vanished and turns and random coiling was seen to be present in 46.6 and 41.0% respectively. This clearly indicates that 12c inhibits β-sheet formation propensity of Aβ42. The spectral curve obtained for Aβ42 (t = 24 h) shows the presence of a positive maxima at 195 nm (black), clearly indicating the conformation of the peptide in the β-sheet form, is absent in the former that is, Aβ42 (t = 0 h), which clearly exhibits a positive maxima at around 205 nm, indicating larger proportion of turn type conformation to be present. No definitive negative minima on the curve were visible at 217 nm, but the shallow curves in the expanded region were indicative of the presence of smaller proportions of α-helix conformations. In the presence of 12c reduction in β-sheet content can also be visualized by the complete absence of the positive maxima at 195 nm, wherein the spectrum follows the similar pattern to that of the Aβ42 (t = 0 h). The prevention of conformational transition to β-sheet suggests the ability of 12c to inhibit the fibrillation process. The positive curve shifts more towards 200–205 nm, indicating a larger proportion of the peptide to be present in the turn form. These observations coincide to the predicted values by the Yang protocol. The effect of the presence of equimolar concentrations of inactive peptide 13a on the conformational changes on Aβ42 was evaluated. A positive maxima at 195 nm is a clear indicative of higher proportions of β-sheet type of secondary structural conformation to be present. This indicates the inactiveness of the peptide 13a and its inability to prevent aggregation of Aβ42. Self-aggregation potential of peptide 13a deters its potential to inhibit Aβ42 aggregation.A compiled CD spectrum recorded depicting a relative comparison of peptides 12f and 13a, incubated for 24 h at 37 °C in the presence and absence of equimolar concentrations of Aβ42 has been presented in Fig. 6. Spectra of test peptides incubated alone were subtracted to obtain the final spectra for comparing the conformational state of Aβ42. A comparison of the individual CD spectrums of the test peptides incubated alone to that incubated in the presence of equimolar ratio of Aβ42 has been summarized (ESI, Fig. S4).Open in a separate windowFig. 6Secondary structure analysis using CD: CD spectrum showing the conformational changes on Aβ42 aggregation in the presence of active peptide 12c and inactive peptide 13a. Aβ42 (10 μM) at 0 h (black) and 24 h (blue), co-incubated individually with equimolar ratios inhibitor peptide 12c (green) and inactive peptide 13a (red) for 24 h.To understand the process of inhibition of Aβ42 in presence of 12c, mass fragmentation techniques were employed.37 The MS spectrum of full-length Aβ42 (10 μM) in the presence and absence of an equimolar amount of 12c was recorded. Fig. 7, shows the ESI spectrum for Aβ42 incubated alone (A), along with 12c (B).Open in a separate windowFig. 7HRMS Analysis: ESI-MS for Aβ42 (10 μM) incubated alone (A), in presence of equimolar ratios of test peptide 12c (B) for 24 h.The spectrum for Aβ42 incubated alone shows a major peak at m/z 685.4357, which may be attributed to aggregated Aβ42 depicting higher mass-to-charge ratio. Further peaks at m/z of 788.4373 [(Aβ42)6+], 507.2713 [(Aβ1–9)1+], 958.3161 [(Aβ10–42)7+] and 1308.0646 [(Aβ1–11)1+] represent the Aβ42 monomer and its specific fragments, respectively. Hexameric form of Aβ42 with m/z 2185.4363 [(6Aβ42)13+] is also observed in the spectrum.38 On comparing the spectrum obtained for Aβ42 incubated along with 12c, and that of Aβ42 incubated alone, additional signal peaks were seen. This suggested the occurrence of adduct between Aβ42 and 12c, as these peaks were not analogous to the molecular weight of native Aβ42 or test peptides themselves. Analyzing the interactions of 12c with that of Aβ42, the mass spectrum shows a peak at m/z 471.7321 [(Aβ12–42 + 12c)8+], depicting 1 : 1 covalent interaction with Aβ12–42 fragment of Aβ42. The spectrum also shows signal peaks corresponding to that of Aβ42, seen in the previous spectrum. It was clear from the above spectrum that the test peptide interacts with the monomeric unit of the Aβ42, thereby preventing its aggregation.Visual investigation of the effects of the peptide 12c, on the morphology and abundance of Aβ42 fibrils was performed by high resolution transmission electron microscopy (HR-TEM).39 Shapes and morphology of the fibrils were also examined using scanning transmission electron microscope (STEM).40 Inactive peptide 13a was selected as a negative control. The control sample of Aβ42 incubated alone at t = 0 h, where uniform distribution of smaller particles of Aβ was seen (Fig. 8A, HR-TEM and Fig. 8D, STEM).Open in a separate windowFig. 8Electron microscopy studies: HR-TEM and STEM images depicting the effects of active peptide 12c and inactive peptide 13a on the aggregation of Aβ42. Aβ42 (10 μM) was incubated alone t = 0 h (A and D), t = 24 h (B and E); with equimolar concentrations of inhibitor peptide 12c (C and F); inactive peptide 13a (H and K) as well as peptide 12c (G and J) and inactive peptide 13a (I and L) incubated alone, respectively. (Additional images have been provided in the ESI, Section 11.3).After an incubation span of 24 h, appearance of amyloid fibrils and an extensive network of long, straw-shaped fibrils were observed (Fig. 8B and E). In the presence of peptide 12c (Fig. 8C and F), only smaller particulate aggregates were seen, indicating complete inhibition of the amyloid fibrils. On co-incubation of Aβ42 with the inactive peptide 13a, large aggregated structures (Fig. 8H and K) were observed. In order to visualize the aggregation of the peptide themselves, equimolar concentrations of peptides 12c and 13a were incubated alone under similar conditions and visualized. Very small granular structures were seen for the 12c (Fig. 8G and J) whereas slightly larger and patchy aggregates were seen for inactive peptide 13a (Fig. 8I and L).In order to evaluate and understand the biosafety and pharmacokinetic profile of the test peptides, cell-cytotoxicity studies employing PC-12 cells was performed. Test peptide were tested up to a highest tested concentration of 20 μM and none of the peptides exhibited undesirable cytotoxicity. Fig. 9A depicts a graphical representation of the % viable cells in presence of 20 μM concentration of peptide 12c.Open in a separate windowFig. 9Cytotoxicity and bioavailability study: (A) analysis of the cytotoxic effects of the peptide 12c (20 mM) on the viability of PC-12 cells evaluated using MTT cell viability assay. The percentage of untreated cells was considered 100% (positive control) and presence of the test peptides in respective dose concentration for 6 h. (B) BBB-permeability of peptide 12c in comparison to 11a, as determined by the PAMPA-BBB assay. Pe was calculated by using the formula VdVa/[(Vd + Va)St] ln(1 − Aa/Ae), where Vd and Va are the mean volumes of the donor and acceptor solutions, S is the surface area of the artificial membrane, t is the incubation time, and Aa and Ae are the UV absorbance of the acceptor well and the theoretical equilibrium absorbance, respectively. Data was recorded for triplicate samples in three individual experiments and the readings were averaged (<5% variation).A major challenge for peptide-based therapeutics is the BBB permeability and proteolytic stability against various enzymes within the body.41In vitro BBB penetration of the most active peptide 12c as well as the lead peptide 11a using parallel artificial membrane permeation assay (PAMPA-BBB) was performed following the previously reported protocols.42–45 The UV/Vis absorptions of both the peptides was recorded after permeating through an artificial porcine polar brain lipid (PBL) membrane and the effective permeabilities (Pe) were calculated. As described in Fig. 9B, the Pe values of the most active peptide 12c was significantly higher than that of 11a, demonstrating enhanced permeability of the modified peptide.Trypsin is the one of the most notorious endopeptidases and cleaves the amide bond next to a charged cationic residue.46 Hence, in order to evaluate whether the synthesized peptides have incorporated the proteolytic stability properties, trypsin and serum stability studies on the peptide 12c was performed. The peptide was incubated with 100-fold excess of trypsin and was subjected to analysis by RP-HPLC. Chromatograms depicted that the peptides exhibited intact integrity, having their retention time unaltered even after 24 h of trypsin treatment. It was observed that, there were no peaks seen before and after the main peak of the peptide indication no fragment and/or other intermediate formation. Superimposed HPLC chromatograms of time point''s intervals have been depicted in Fig. 10A.Open in a separate windowFig. 10Proteolytic stability study: (A) superimposed HPLC chromatograms of most active peptide 12c at time intervals of 0, 2, 4, 8, 12, 18 and 24 h after trypsin treatment; (B) graphical representation showing % degradation for peptide 12c on serum treatment. (C) Mass spectra for peptide 12c at 0, 12, 18 and 24 h of serum treatment. Analyzed by ACD-Mass Fragmenter tool. (D) Predicted susceptible cleavage sites for peptide 12c. Most susceptible peptide bond has been indicated in bold red.Serum stability assay following similar protocol was performed. The chromatograms depicted that the peptides exhibited intact integrity, having their retention time unaltered up to 12 h of serum treatment. The peaks obtained for the peptides at 18 h and 24 h of serum treatment, were comparatively smaller to that of the previous peaks indicating slight degradation of the peptide. Superimposed HPLC chromatogram for peptide 12c has been provided in ESI, Fig. S6.To calculate the rate of degradation of the peptide on serum treatment, analysis and comparison the area under the curve of the peak of the respective peptide at their specific time intervals.47,48 A comparative analysis depicted that around 20% of the peptide is present after 24 h of serum treatment. Fig. 10B indicates the % degradation of the peptide in serum over a period of 24 h. The initial calculation of % peptide present in the sample aliquot is indicated in the ESI, Fig. S7. The extrapolated data helped us to determine the degradation rate of the peptide in serum (ESI, Table S8).48–50 It can be concluded that approximately 50% of the peptide is stable until 12 h of serum treatment, following which it shows an decline in stability decreasing to about 22% until 24 h.In order to study the mode of degradation of the peptide and the susceptibility of the peptide towards peptide degradation, mass spectroscopy was employed. The samples analyzed for peptide content on RP-HPLC were further subjected to LCQ analysis.51,52 At 0 h, the molecular ion peak (m/z 470) of the peptide corresponded to the molecular mass of the peptide itself. Sequential fragmentation pattern was minimal and a clear mass spectrum was seen. After 12 h, multiple fragmentation peaks were seen indicating that the peptide has undergone cleavage at multiple sites. ACD-Mass Fragmenter tool was used to analyze the specific fragmentation pattern. Fig. 10C summarizes the mass spectrums and shows the specific fragmentation pattern. Based on the fragmentation pattern for the tetrapeptide sequence, the most susceptible bonds that could easily undergo cleavage were identified. ACD Mass Fragmenter tool was used to understand the peptide fragmentation pattern. Fig. 10D depicts the structure of the peptide 12c indicating the most susceptible peptide bond. This understanding provides impetus in site specific modification in improving the stability of the peptides.Computationally understanding the binding mechanism and intra-residual interactions of the test peptides with the single monomeric as well as the proto-fibrillar unit of Aβ42 would prove to be useful. Various structures for the both the forms of Aβ42 are reported in the literature. A monomeric sequence bearing complete sequence of all 42 amino acids residues was used (PDB Id: 1IYT; ESI, Fig. S8). The structure comprises of α-helices and random coiling, similar to the data previously reported.53 A proto-fibrillar unit comprising of 6 full length spatially arranged in a ‘S’ shaped manner recently reported by Colvin and co-workers54 (PDB Id: 2NAO, ESI, Fig. S10) was used for understanding the interaction of the test peptides on the proto-fibrillar unit of Aβ. Reactive site analysis of the pre-optimized framework of the monomeric Aβ42 showed that the hinge region is the most prone to aggregation due to the presence of reactive residues in that particular segment (Maestro, Bioluminate suite; ESI Fig. S9). To understand the interaction of the test peptide with the full-length Aβ42, a grid incorporating the whole sequence was generated. Docking studies were performed with the peptides mentioned in this work along with a few molecules reported in literature10–12,55,56 for comparative analysis of the binding modes and interactions.57–59 Docking, glide and residual interaction energy scores for the molecules selected from literature and test peptides from the current study are summarized in ESI (Tables S9 and S10). Although the analysis of the scores reveal that similar docking and glide scores were seen in both the cases. The energy of interaction of the test peptides with the monomeric unit proved to be a comparable factor to that of the literature reported ligands (ESI Table S11).Test peptides have shown to interact with Glu11, His13, His14, Gln15, Phe19, Phe20 and more specifically with Asp23. These residues are involved to aid the aggregation of monomeric Aβ42 into the fibrillar species, as also seen in reactive residue analysis (ESI, Fig. S11). Binding of the test peptides to these residues, block the free interaction of these residues to others, inhibiting their aggregation propensity. Ligand interactions of the most active test peptide 12c with the monomeric unit, their respective ligand interaction diagrams (2D and 3D) have been summarized in Fig. 11A and B. The interaction energies are in accordance to those exhibited by the standard ligands indicating similar kind of interactions and thus reinforcing the results of studies carried out in this work. A structural framework 2NAO-06 of proto-fibrillar Aβ42 was rationalized to be the most optimal structure and was prepared for docking studies (ESI, Fig. S10). Since it is a proto-fibrillar unit, the most reactive sites for the binding were identified. SiteMap Analysis feature identified 5 ligand-binding sites (ESI, Fig. S12). The predicted sites did coincide with the predicted reactive residues, thus indicating certain interactions between those residues and the ligand to be feasible, which would inhibit the process of aggregation. To carry out docking studies, receptor grids at the predicted sites on the proto-fibrillar unit was generated. Since there has been no docking studies performed using 2NAO as the protein, validation of the use of 2NAO and the predicted sites, by docking standard ligands was performed (ESI, Tables S12 and S13). Analysis of the docking studies revealed SiteMap-2 to be the most plausible site for action of an inhibitor. The molecule can interact with the residues that aid in aggregation. This would block the further attachment of another monomeric unit to the site-recognition units on the proto-fibrillar structure. Subsequent interaction with the neighboring residues of both the chains destabilizes the preformed bonds, which hold the adjacent units together.Open in a separate windowFig. 11 In silico study: ligand interaction diagram showing interactions of the ligand with the residues of monomeric unit 1IYT-10 (A and B) and with the proto-fibrillar unit 2NAO-06 (C and D). 3D representation (Left) showed along with 2D representation (Right).Docking studies of the peptides presented in this work was carried out and docking scores and the residue interaction energies obtained for the set of synthesized tetrapeptides for SiteMap-2 has been summarized in ESI, Table S14. On careful analysis it can be seen that interaction with Met35, Val36, Gly37 of chain D, as well as Glu11. His13, His14 and Gln15 of the neighboring chain A, show that the test peptide does interact with both the neighboring chains and especially with those amino acids that are solely responsible for maintaining the dimeric structure of the proto-fibrillar unit. To have a clear understanding of the interactions, ligand interaction diagrams for the test peptide 12c, depicting its interaction with the specific residues of the proto-fibrillar unit have been summarized in Fig. 11C and D.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号