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 203 毫秒
1.
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   

2.
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   

3.
Research continues to find a breakthrough for the treatment of Alzheimer''s Disease (AD) due to its complicated pathology. Presented herein is a novel series of arydiazoquinoline molecules investigated for their multifunctional properties against the factors contributing to Alzheimer''s disease (AD). The inhibitory properties of fourteen closely related aryldiazoquinoline derivatives have been evaluated for their inhibitory effect on Aβ42 peptide aggregation. Most of these molecules inhibited Aβ42 fibrillation by 50–80%. Selected molecules were also investigated for their binding behaviour to preformed Aβ40 aggregates indicating a nanomolar affinity. In addition, these compounds were further investigated as cholinesterase inhibitors. Interestingly, some of the compounds turned out to be moderate in vitro inhibitors for AChE activity with IC50 values in low micro molar range. The highest anti-AChE activity was shown by compound labelled as 2a with an IC50 value of 6.2 μM followed by 2b with IC50 value of 7.0 μM. In order to understand the inhibitory effect, binding of selected molecules to AChE enzyme was studied using molecular docking. In addition, cell toxicity studies using Neuro2a cells were performed to assess their effect on neuronal cell viability which suggests that these molecules possess a non-toxic molecular framework. Overall, the study identifies a family of molecules that show good in vitro anti-Aβ-aggregation properties and moderately inhibit cholinesterase activity.

Novel series of aryldiazoquinoline multifunctional molecules controls amyloid formation and neuro-protective role by inhibiting esterase enzymes.  相似文献   

4.
A novel series of benzothiazole–piperazine hybrids were rationally designed, synthesized, and evaluated as multifunctional ligands against Alzheimer''s disease (AD). The synthesized hybrid molecules illustrated modest to strong inhibition of acetylcholinesterase (AChE) and Aβ1-42 aggregation. Compound 12 emerged as the most potent hybrid molecule exhibiting balanced functions with effective, uncompetitive and selective inhibition against AChE (IC50 = 2.31 μM), good copper chelation, Aβ1-42 aggregation inhibition (53.30%) and disaggregation activities. Confocal laser scanning microscopy and TEM analysis also validate the Aβ fibril inhibition ability of this compound. Furthermore, this compound has also shown low toxicity and is capable of impeding loss of cell viability elicited by H2O2 neurotoxicity in SHSY-5Y cells. Notably, compound 12 significantly improved cognition and spatial memory against scopolamine-induced memory deficit in a mouse model. Hence, our results corroborate the multifunctional nature of novel hybrid molecule 12 against AD and it may be a suitable lead for further development as an effective therapeutic agent for therapy in the future.

A novel series of benzothiazole–piperazine hybrids were rationally designed, synthesized, and evaluated as multifunctional ligands against Alzheimer''s disease (AD).  相似文献   

5.
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   

6.
Aberrant misfolding and amyloid aggregation, which result in amyloid fibrils, are frequent and critical pathological incidents in various neurodegenerative disorders. Multiple drugs or inhibitors have been investigated to avert amyloid aggregation in individual peptides, exhibiting sequence-dependent inhibition mechanisms. Establishing or inventing inhibitors capable of preventing amyloid aggregation in a wide variety of amyloid peptides is quite a daunting task. Bleomycin (BLM), a complex glycopeptide, has been widely used as an antibiotic and antitumor drug due to its ability to inhibit DNA metabolism, and as an antineoplastic, especially for solid tumors. In this study, we investigated the dual inhibitory effects of BLM on Aβ aggregation, associated with Alzheimer''s disease and hIAPP, which is linked to type 2 diabetes, using both computational and experimental techniques. Combined results from drug repurposing and replica exchange molecular dynamics simulations demonstrate that BLM binds to the β-sheet region considered a hotspot for amyloid fibrils of Aβ and hIAPP. BLM was also found to be involved in β-sheet destabilization and, ultimately, in its reduction. Further, experimental validation through in vitro amyloid aggregation assays was obtained wherein the fibrillar load was decreased for the BLM-treated Aβ and hIAPP peptides in comparison to controls. For the first time, this study shows that BLM is a dual inhibitor of Aβ and hIAPP amyloid aggregation. In the future, the conformational optimization and processing of BLM may help develop various efficient sequence-dependent inhibitors against amyloid aggregation in various amyloid peptides.

Bleomycin acts as a dual inhibitor against both amyloid β and human islet amyloid polypeptide by binding to the β-sheet grooves considered as the amyloids hotspot.  相似文献   

7.
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   

8.
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   

9.
Atomistic investigation of an Iowa Amyloid-β trimer in aqueous solution     
Son Tung Ngo  Huong Thi Thu Phung  Khanh B. Vu  Van V. Vu 《RSC advances》2018,8(73):41705
The self-assembly of Amyloid beta (Aβ) peptides are widely accepted to associate with Alzheimer''s disease (AD) via several proposed mechanisms. Because Aβ oligomers exist in a complicated environment consisting of various forms of Aβ, including oligomers, protofibrils, and fibrils, their structure has not been well understood. The negatively charged residue D23 is one of the critical residues of the Aβ peptide as it is located in the central hydrophobic domain of the Aβ N-terminal and forms a salt-bridge D23-K28, which helps stabilize the loop domain. In the familial Iowa (D23N) mutant, the total net charge of Aβ oligomers decreases, resulting in the decrease of electrostatic repulsion between D23N Aβ monomers and thus the increase in their self-aggregation rate. In this work, the impact of the D23N mutation on 3Aβ11–40 trimer was characterized utilizing temperature replica exchange molecular dynamics (REMD) simulations. Our simulation reveals that D23N mutation significantly enhances the affinity between the constituting chains in the trimer, increases the β-content (especially in the sequence 21–23), and shifts the β-strand hydrophobic core from crossing arrangement to parallel arrangement, which is consistent with the increase in self-aggregation rate. Molecular docking indicates that the Aβ fibril-binding ligands bind to the D23N and WT forms at different poses. These compounds prefer to bind to the N-terminal β-strand of the D23N mutant trimer, while they mostly bind to the N-terminal loop region of the WT. It is important to take into account the difference in the binding of ligands to mutant and wild type Aβ peptides in designing efficient inhibitors for various types of AD.

Amyloid beta peptide oligomers are believed to play key roles in Alzheimer''s disease pathogenesis. D23N mutation significantly changes their structure and how they bind potential inhibitors.  相似文献   

10.
Decarboxylative aldol reaction of α,α-difluoro-β-ketocarboxylate salt: a facile method for generation of difluoroenolate     
Atsushi Tarui  Mayuna Oduti  Susumu Shinya  Kazuyuki Sato  Masaaki Omote 《RSC advances》2018,8(37):20568
We developed a decarboxylative aldol reaction using α,α-difluoro-β-ketocarboxylate salt, carbonyl compounds, and ZnCl2/N,N,N′,N′-tetramethylethylenediamine. The generation of difluoroenolate proceeded smoothly under mild heating to provide α,α-difluoro-β-hydroxy ketones in good to excellent yield (up to 99%). The α,α-difluoro-β-ketocarboxylate salt was bench stable and easy to handle under air, which realizes a convenient and environmentally friendly methodology for synthesis of difluoromethylene compounds.

A ZnCl2/N,N,N′,N′-tetramethylethylenediamine complex promoted decarboxylative aldol reaction of α,α-difluoro-β-ketocarboxylate salt with carbonyl compounds has been developed.  相似文献   

11.
Development of an MRI contrast agent for both detection and inhibition of the amyloid-β fibrillation process     
Rohmad Yudi Utomo  Satoshi Okada  Akira Sumiyoshi  Ichio Aoki  Hiroyuki Nakamura 《RSC advances》2022,12(8):5027
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   

12.
A Maitake (Grifola frondosa) polysaccharide ameliorates Alzheimer's disease-like pathology and cognitive impairments by enhancing microglial amyloid-β clearance     
Yao Bai  Lingling Chen  Yao Chen  Xinmeng Chen  Yilong Dong  Shangyong Zheng  Lei Zhang  Weiyuan Li  Jing Du  Hongliang Li 《RSC advances》2019,9(64):37127
Alzheimer''s disease (AD) is characterized by the deposition of amyloid-β (Aβ) plaques, neuronal loss and neurofibrillary tangles. In addition, neuroinflammatory processes are thought to contribute to AD pathophysiology. Maitake (Grifola frondosa), an edible/medicinal mushroom, exhibits high nutritional value and contains a great amount of health-beneficial, bioactive compounds. It has been reported that proteo-β-glucan, a polysaccharide derived from Maitake (PGM), possesses strong immunomodulatory activities. However, whether PGM is responsible for the immunomodulatory and neuroprotection effects on APPswe/PS1ΔE9 (APP/PS1) transgenic mice, a widely used animal model of AD, remains unclear. In the present study, the results demonstrated that PGM could improve learning and memory impairment, attenuate neuron loss and histopathological abnormalities in APP/PS1 mice. In addition, PGM treatment could activate microglia and astrocytes and promote microglial recruitment to the Aβ plaques. Also, PGM could enhance Aβ phagocytosis, and thereby alleviate Aβ burden and the pathological changes in the cortex and hippocampus in APP/PS1 mice. Moreover, PGM showed no significant effect on mice body weight. In conclusion, these findings indicated that administration of PGM could improve memory impairment via immunomodulatory action, and dietary supplementation with PGM may provide potential benefits on brain aging related memory dysfunction.

PGM ameliorates AD-like pathology and cognitive impairments by enhancing microglial amyloid-β clearance.  相似文献   

13.
Synthesis of 1-(β-coumarinyl)-1-(β-indolyl)trifluoroethanols through regioselective Friedel–Crafts alkylation of indoles with β-(trifluoroacetyl)coumarins catalyzed by Sc(OTf)3     
Lijun Shi  Ying Liu  Caixia Wang  Xinxin Yuan  Xiaobiao Liu  Lulu Wu  Zhenliang Pan  Qicheng Yu  Cuilian Xu  Guoyu Yang 《RSC advances》2020,10(24):13929
A highly efficient Friedel–Crafts alkylation of indole derivatives with β-(trifluoroacetyl)coumarins using Sc(OTf)3 as a catalyst has been developed, which gives regioselective 1,2-adducts to afford 1-(β-coumarinyl)-1-(β-indolyl)trifluoroethanols. A series of tertiary trifluoroethanols containing different indole and coumarin groups were synthesized in moderate to excellent yields (up to 95%) in the presence of 5 mol% catalyst in a short time (only 2 minutes at least). A mechanism of the reaction, in which the trace amount of water plays the role of proton transfer in catalyzing circulation was proposed and confirmed.

A Friedel–Crafts alkylation of indoles with β-(trifluoroacetyl)coumarins catalyzed by Sc(OTf)3 to afford 1-(β-coumarinyl)-1-(β-indolyl)trifluoroethanols in a short time and high yield was developed.  相似文献   

14.
Gene therapy using Aβ variants for amyloid reduction     
Kyung-Won Park  Caleb A. Wood  Jun Li  Bethany C. Taylor  SaeWoong Oh  Nicolas L. Young  Joanna L. Jankowsky 《Molecular therapy》2021,29(7):2294
Numerous aggregation inhibitors have been developed with the goal of blocking or reversing toxic amyloid formation in vivo. Previous studies have used short peptide inhibitors targeting different amyloid β (Aβ) amyloidogenic regions to prevent aggregation. Despite the specificity that can be achieved by peptide inhibitors, translation of these strategies has been thwarted by two key obstacles: rapid proteolytic degradation in the bloodstream and poor transfer across the blood-brain barrier. To circumvent these problems, we have created a minigene to express full-length Aβ variants in the mouse brain. We identify two variants, F20P and F19D/L34P, that display four key properties required for therapeutic use: neither peptide aggregates on its own, both inhibit aggregation of wild-type Aβ in vitro, promote disassembly of pre-formed fibrils, and diminish toxicity of Aβ oligomers. We used intraventricular injection of adeno-associated virus (AAV) to express each variant in APP/PS1 transgenic mice. Lifelong expression of F20P, but not F19D/L34P, diminished Aβ levels, plaque burden, and plaque-associated neuroinflammation. Our findings suggest that AAV delivery of Aβ variants may offer a novel therapeutic strategy for Alzheimer’s disease. More broadly our work offers a framework for identifying and delivering peptide inhibitors tailored to other protein-misfolding diseases.  相似文献   

15.
Catalyst-free chemoselective α-sulfenylation/β-thiolation for α,β-unsaturated carbonyl compounds     
Xi Huang  Juan Li  Xiang Li  Jiayi Wang  Yanqing Peng  Gonghua Song 《RSC advances》2019,9(45):26419
A novel, efficient, catalyst-free and product-controllable strategy has been developed for the chemoselective α-sulfenylation/β-thiolation of α,β-unsaturated carbonyl compounds. An aromatic sulfur group could be chemoselectively introduced at α- or β-position of carbonyls with different sulfur reagents under slightly changed reaction conditions. A series of desired products were obtained in moderate to excellent yields. Mechanistic studies revealed that B2pin2 played the key role in activating the transformation towards the β-thiolation of α,β-unsaturated carbonyl compounds. This transition-metal-catalyst-free method provides a convenient and efficient tool for the highly chemoselective preparation of α-thiolation or β-sulfenylation products of α,β-unsaturated carbonyl compounds.

This catalyst-free method provides a useful and efficient tool for the highly chemoselective preparation of α-thiolation or β-sulfenylation products of α,β-unsaturated carbonyl compounds.  相似文献   

16.
New α-pyrones from an endophytic fungus,Hypoxylon investiens J2     
Chao Yuan  Hong-Xia Yang  Yu-Hua Guo  Lin Fan  Ying-Bo Zhang  Gang Li 《RSC advances》2019,9(47):27419
Four new α-pyrones, hypotiens A–D (1–4), were isolated from a fungal endophyte, Hypoxylon investiens J2, harbored in the medicinal plant Blumea balsamifera. Their structures were determined through detailed HRMS and NMR spectroscopic data. Compounds 1–4 are new α-pyrone derivatives containing an unusual dimethyl substitution in the highly unsaturated side chain. Their plausible biosynthetic pathway was discussed. Biological assay indicated that compounds 1–4 showed no antimicrobial, quorum sensing inhibitory, and cytotoxic activities. The specific side chain in α-pyrone derivatives 1–4 might be responsible for the weak pharmacological activities.

Four new α-pyrones, hypotiens A–D (1–4), were isolated from a fungal endophyte, Hypoxylon investiens J2, harbored in the medicinal plant Blumea balsamifera.  相似文献   

17.
Geniposide attenuates Aβ25–35-induced neurotoxicity via the TLR4/NF-κB pathway in HT22 cells     
Xiu-Fang Huang  Jian-Jun Li  Yan-Gu Tao  Xie-Qi Wang  Ru-Lan Zhang  Jia-Lin Zhang  Zu-Qing Su  Qi-Hui Huang  Yuan-Hui Deng 《RSC advances》2018,8(34):18926
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   

18.
Direct β-selectivity of α,β-unsaturated γ-butyrolactam for asymmetric conjugate additions in an organocatalytic manner     
Yuan Zhong  Sihua Hong  Zhengjun Cai  Shixiong Ma  Xianxing Jiang 《RSC advances》2018,8(51):28874
The β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by a bifunctional chiral tertiary amine has been developed, which provides an efficient access to optically active β-position functionalized pyrrolidin-2-one derivatives in both high yield and enantioselectivity (up to 78% yield and 95 : 5 er). This is the first catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach.

The asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by (DHQD)2AQN has been developed, which provides an access to β-position functionalized pyrrolidin-2-one derivatives in high levels yield and enantioselectivity.

Metal-free organocatalytic asymmetric transformations have successfully captured considerable enthusiasm of chemists as powerful methods for the synthesis of various kinds of useful chiral compounds ranging from the preparation of biologically important molecules through to novel materials.1 Chiral pyrrolidin-2-ones have been recognized as important structural motifs that are frequently encountered in a variety of biologically active natural and synthetic compounds.2 In particular, the β-position functionalized pyrrolidin-2-one backbones, which can serve as key synthetic precursors for inhibitory neurotransmitters γ-aminobutyric acids (GABA),3 selective GABAB receptor agonists4 as well as antidepressant rolipram analogues,5 have attracted a great deal of attention. Therefore, the development of highly efficient, environmentally friendly and convenient asymmetric synthetic methods to access these versatile frameworks is particularly appealing.As a direct precursor to pyrrolidin-2-one derivatives, recently, α,β-unsaturated γ-butyrolactam has emerged as the most attractive reactant in asymmetric organometallic or organocatalytic reactions for the synthesis of chiral γ-position functionalized pyrrolidin-2-ones (Scheme 1). These elegant developments have been achieved in the research area of catalytic asymmetric vinylogous aldol,6 Mannich,7 Michael8 and annulation reactions9 in the presence of either metal catalysts or organocatalysts (a, Scheme 1). These well-developed catalytic asymmetric methods have been related to the γ-functionalized α,β-unsaturated γ-butyrolactam to date. However, in sharp contrast, the approaches toward introducing C-3 chirality at the β-position of butyrolactam through a direct catalytic manner are underdeveloped (b, Scheme 1)10 in spite of the fact that β-selective chiral functionalization of butyrolactam can directly build up α,β-functionalized pyrrolidin-2-one frameworks.Open in a separate windowScheme 1Different reactive position of α,β-unsaturated γ-butyrolactam in catalytic asymmetric reactions.So far, only a few metal-catalytic enantioselective β-selective functionalized reactions have been reported. For examples, a rhodium/diene complex catalyzed efficient asymmetric β-selective arylation10a and alkenylation10b have been reported by Lin group (a, Scheme 2). Procter and co-workers reported an efficient Cu(i)–NHC-catalyzed asymmetric silylation of unsaturated lactams (b, Scheme 2).10c Despite these creative works, considerable challenges still exist in the catalytic asymmetric β-selective functionalization of γ-butyrolactam. First, the scope of nucleophiles is limited to arylboronic acids, potassium alkenyltrifluoroborates and PhMe2SiBpin reagents. Second, the catalytic system and activation mode is restricted to metal/chiral ligands. To our knowledge, an efficient catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach has not yet been established. Therefore, the development of organocatalytic asymmetric β-selective functionalization of γ-butyrolactam are highly desirable. In conjunction with our continuing efforts in building upon chiral precedents by using chiral tertiary amine catalytic system,11 we rationalized that the activated α,β-unsaturated γ-butyrolactam might serve as a β-position electron-deficient electrophile. This γ-butyrolactam may react with a properly designed electron-rich nucleophile to conduct an expected β-selective functionalized reaction of γ-butyrolactam under a bifunctional organocatalytic fashion, while avoiding the direct γ-selective vinylogous addition reaction or β,γ-selective annulation as outlined in Scheme 2. Herein we report the β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters12 catalyzed by a bifunctional chiral tertiary amine, which provides an efficient and facile access to optically active β-position functionalized pyrrolidin-2-one derivatives with both high diastereoselectivity and enantioselectivity.Open in a separate windowScheme 2β-Selective functionalization of γ-butyrolactam via metal- (previous work) or organo- (this work) catalytic approach.To begin our initial investigation, several bifunctional organocatalysts13 were firstly screened to evaluate their ability to promote the β-selective asymmetric addition of γ-butyrolactam 2a with cyclic imino ester 3a in the presence of 15 mol% of catalyst loading at room temperature in CH2Cl2 (entries 1–6, EntryCat.SolventYieldeerf11aCH2Cl270%40 : 6021bCH2Cl2<5%57 : 4331cCH2Cl270%65 : 3541dCH2Cl268%70 : 3051eCH2Cl258%63 : 4761fCH2Cl271%77 : 2371fDCE72%80 : 2081fCHCl370%80 : 2091fMTBE68%79 : 21101fToluene63%78 : 22111fTHF45%76 : 24121fMeOH32%62 : 3813b1fDCE : MTBE75%87 : 1314c1fDCE : MTBE72%87 : 1315d1fDCE : MTBE70%85 : 15Open in a separate windowaReaction conditions: unless specified, a mixture of 2a (0.2 mmol), 3a (0.3 mmol) and a catalyst (15 mmol%) in a solvent (2.0 mL) was stirred at rt. for 48 h.bThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1).cThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) for 24 h.dThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) and 10 mol% of catalyst was used.eIsolated yields.fDetermined by chiral HPLC, the product was observed with >99 : 1 dr by 1H NMR and HPLC. Configuration was assigned by X-ray crystal data of 4a.The results of experiments under the optimized conditions that probed the scope of the reaction are summarized in Scheme 3. The catalytic β-selective asymmetric addition of γ-butyrolactam 2a with cyclic imino esters 3a in the presence of 15 mol% (DHQD)2AQN 1f was performed. A variety of phenyl-substituted cyclic imino esters including those bearing electron-withdrawing and electron-donating substituents on the aryl ring, heterocyclic were also examined. The electron-neutral, electron-rich, or electron-deficient groups on the para-position of phenyl ring of the cyclic imino esters afforded the products 4a–4m in 57–75% yields and 82 : 18 to 95 : 5 er values. It appears that either an electron-withdrawing or an electron-donating at the meta- or ortho-position of the aromatic ring had little influence on the yield and stereoselectivity. Similar results on the yield and enantioselectivities were obtained with 3,5-dimethoxyl substituted cyclic imino ester (71% yield and 91 : 9 er). It was notable that the system also demonstrated a good tolerance to naphthyl substituted imino ester (78% yield and 92 : 8 er value). The 2-thienyl substituted cyclic imino ester proceeded smoothly under standard conditions as well, which gave the desired product 4p in good enantioselectivity (88 : 12 er), although yield was slightly lower. However, attempts to extend this methodology to aliphatic-substituted product proved unsuccessful due to the low reactivity of the substrate 3q. It is worth noting that the replacement of Boc group with 9-fluorenylmethyl, tosyl or benzyl group as the protection, no reaction occurred. The absolute and relative configurations of the products were unambiguously determined by X-ray crystallography (4a, see the ESI).Open in a separate windowScheme 3Substrate scope of the asymmetric reaction of α,β-unsaturated γ-butyrolactam 2 to cyclic imino esters 3.a aReaction conditions: unless specified, a mixture of 2 (0.2 mmol), 3 (0.3 mmol) and 1f (15.0 mmol%) in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) was stirred at rt. bIsolated yields. cDetermined by chiral HPLC, all products were observed with >99 : 1 dr by 1H NMR and HPLC. Configuration was assigned by comparison of HPLC data and X-ray crystal data of 4a.We then examined the substrate scope of the imide derivatives (Scheme 4). Investigations with maleimides 4r–4u gave 48–61% yield of corresponding products as lower er and dr values than most of γ-butyrolactams. As for methyl substituted maleimides, the reaction failed to give any product.Open in a separate windowScheme 4Substrate scope of the asymmetric reaction of maleimides to cyclic imino esters.a aReaction conditions: unless specified, a mixture of 2 (0.2 mmol), 3 (0.3 mmol) and 1f (15.0 mmol%) in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) was stirred at rt. bIsolated yields. cDetermined by 1H NMR and chiral HPLC.The chloride product 4a ((R)-tert-butyl 4-((R)-3-((E)-(4-chlorobenzylidene)amino)-2-oxotetra hydrofuran-3-yl)-2-oxopyrrolidine-1-carboxylate) was recrystallized and the corresponding single crystal was subjected to X-ray analysis to determine the absolute structure. Based on this result and our previous work, a plausible catalytic mechanism involving multisite interactions was assumed to explain the high stereoselectivity of this process (Fig. 1). Similar to the conformation reported for the dihydroxylation and the asymmetric direct aldol reaction, the transition state structure of the substrate/catalyst complexes might be presumably in the open conformation. The acidic α-carbon atom of cyclic imino ester 3a could be activated by interaction between the tertiary amine moiety of the catalyst and the enol of 3avia a hydrogen bonding. Moreover, the enolate of 3a in the transition state might be in part stabilized through the π–π stacking between the phenyl ring of 3a and the quinoline moiety. Consequently, the Re-face of the enolate is blocked by the left half of the quinidine moiety. The steric hindrance between the Boc group of 2a and the right half of the quinidine moiety make the Re-face of 2a face to the enolate of 3a. Subsequently, the attack of the incoming nucleophiles forms the Si-face of enolate of 3a to Re-face of 2a takes place, which is consistent with the experimental results.Open in a separate windowFig. 1Proposed transition state for the reaction.In conclusion, we have disclosed the β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by a bifunctional chiral tertiary amine, which provides an efficient and facile access to optically active β-position functionalized pyrrolidin-2-one derivatives with high diastereoselectivity and enantioselectivity. To our knowledge, this is the first catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach. Current efforts are in progress to apply this new methodology to synthesize biologically active products.  相似文献   

19.
N α-arylsulfonyl histamines as selective β-glucosidase inhibitors     
M. O. Salazar  M. I. Osella  I. A. Ramallo  R. L. E. Furlan 《RSC advances》2018,8(63):36209
N α-benzenesulfonylhistamine, a new semi-synthetic β-glucosidase inhibitor, was obtained by bioactivity-guided isolation from a chemically engineered extract of Urtica urens L. prepared by reaction with benzenesulfonyl chloride. In order to identify better β-glucosidase inhibitors, a new series of Nα,Nτ-di-arylsulfonyl and Nα-arylsulfonyl histamine derivatives was prepared. Biological studies revealed that the β-glucosidase inhibition was in a micromolar range for several Nα-arylsulfonyl histamine compounds of the series, Nα-4-fluorobenzenesulfonyl histamine being the most powerful compound. Besides, this reversible and competitive inhibitor presented a good selectivity for β-glucosidase with respect to other target enzymes including α-glucosidase.

A selective β-glucosidase inhibitor was discovered using the chemically engineered extracts approach.  相似文献   

20.
Structural and electronic properties of α-, β-, γ-, and 6,6,18-graphdiyne sheets and nanotubes     
Linwei Li  Weiye Qiao  Hongcun Bai  Yuanhe Huang 《RSC advances》2020,10(28):16709
α-, β-, γ- and 6,6,18-graphdiyne (GDYs) sheets, as well as the corresponding nanotubes (GDYNTs) are investigated systematically by using the self-consistent-field crystal orbital method. The calculations show that the GDYs and GDYNTs with different structures have different electronic properties. The α-GDY sheet is a conductor, while 2D β-, γ- and 6,6,18-GDYs are semiconductors. The carrier mobilities of β- and γ-GDY sheets in different directions are almost the same, indicating the isotropic transport characteristics. In addition, the electron mobility is in the order of 106 cm2 V−1 s−1 and it is two orders of magnitude larger than the hole mobility of 2D γ-GDY. However, α- and 6,6,18-GDY sheets have anisotropic mobilities, which are different along different directions. For the 1D tubes, the order of stability is γ-GDYNTs > 6,6,18-GDYNTs > β-GDYNTs > α-GDYNTs and is independent of the tube chirality and size. β- and γ-GDYNTs as well as zigzag α- and 6,6,18-GDYNTs are semiconductors with direct bandgaps, while armchair α-GDYNTs are metals, and armchair 6,6,18-GDYNTs change from semiconductors to metals with increasing tube size. The armchair β- and γ-GDYNTs are more favourable to transport holes, while the corresponding zigzag tubes prefer to transport electrons.

Theoretical investigation of α-, β-, γ- and 6,6,18-graphdiyne sheets as well as their corresponding nanotubes.  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