Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Bis(sulfosuccinimidyl) suberate (BS3) crosslinking analysis of the behavior of amyloid-β peptide in solution and in phospholipid membranes

  • Jing-Ming Shi,

    Affiliations Research Institute of Atherosclerotic Disease, Xi’an Jiaotong University Cardiovascular Research Center, Xi’an P. R. China, School of Medicine, Xizang Minzu University, Xian’ yang P. R. China

  • Jie Pei,

    Affiliation Lanzhou Institute of Husbandry and Pharmaceutical Sciences, Chinese Academy of Agricultural Sciences, Lanzhou P. R. China

  • En-Qi Liu ,

    liuenqi@mail.xjtu.edu.cn

    Affiliation Research Institute of Atherosclerotic Disease, Xi’an Jiaotong University Cardiovascular Research Center, Xi’an P. R. China

  • Lin Zhang

    Affiliation Research Institute of Atherosclerotic Disease, Xi’an Jiaotong University Cardiovascular Research Center, Xi’an P. R. China

Abstract

The structure and state of amyloid-β peptide (Aβ) oligomers often need to be checked by reliable experimental methods. Electrophoresis is a commonly applied measurement method. However, due to the presence of detergents, oligomers are easily broken during electrophoresis, which makes it very hard to accurately assess Aβ aggregate states. In the current study, bis(sulfosuccinimidyl) suberate (BS3) was used to cross-link Aβ1–42 oligomers prior to electrophoresis. When compared to a previously reported Aβ cross-linking agent, glutaraldehyde, it was quite apparent that BS3 is more suitable for detecting intra-membrane Aβ oligomers and extra-membrane Aβ oligomers states. As such, our findings provide an efficient method for analyzing Aβ proteins or other proteins that are easily aggregated in solution and in phospholipid membranes.

Introduction

Alzheimer’s disease (AD) is the most common cause of dementia. Worldwide, there are 40 million people who have dementia, including 60–70% of AD cases, and this number is expected to double every 20 years [1]. The pathological changes in AD brain tissues include the extracellular deposition of neuritic plaques that are mainly composed of amyloid-β peptide(Aβ), as well as the intracellular formation of neurofibrillary tangles induced by hyperphosphorylated tau (p-tau). Although the pathogenesis of AD is unclear, after several decades of research, Aβ is still considered the main pathogenic factor for the disease [2,3]. Aβ in cerebrospinal fluid, total tau protein content and phosphorylated tau content have been used as biological markers and detected in clinical trials [4,5], and these indexes are indispensable in the prevention and treatment of AD.

Aβ is generated from amyloid precursor protein via sequential hydrolysis by β- and γ-secretases and rapidly forms soluble oligomers upon release. Aβ oligomers play critical toxic roles in the initial stage of AD [6]. These oligomers are not comprised of a single oligomer, but include various different oligomers.This is also the case for homogeneous Aβ-derived diffusible ligands (ADDL), which, by electron microscopy have been found to contain a mixture of oligomers ranging from 3-mers to 24-mers [7]. As such, it is difficult to determine which oligomers are responsible for Aβ toxicity. Therefore, assessing the ability of Aβ to aggregate is necessary to understand the mechanisms underlying its toxicity.

Electrophoresis is a frequently used method for detecting Aβ aggregates. During electrophoresis, the Aβ1–42 oligomer structure is damaged by the denaturing agents, such as sodium dodecyl sulfonate (SDS), and physically constrained by polyacrylamide gel electrophoresis (PAGE) channels. On electrophoresis gels, soluble Aβ oligomers usually display bands that correspond to three forms, monomers, trimers and tetramers [8]. Size exclusion chromatography (SEC) has been used to isolate low-molecular-weight Aβ oligomers and high-molecular-weight Aβ aggregates; moreover, SEC is one of the most commonly used methods for isolating and purifying Aβ oligomers[9,10], while its original aggregated form is unknown

Crosslinking helps to stabilize the structure of oligomers and can thus be used to reflect their native structure. Glutaraldehyde and photo-crosslinking have been used to analyze oligomers [1113].

These experiments focused more on the results of cross-linking but did not pay attention to whether they were crosslinking monomers into oligomers. If the monomers are cross-linked into oligomers, the reaction is obviously overdone, and does not achieve the purpose of cross-linking. To achieve higher resolution, we tested the efficacy of bis(sulfosuccinimidyl)(BS3), a water-soluble compound that has previously been used in the cross-linking of antibodies [14,15]. BS3 has previously been used to cross-link intracellular epidermal growth factor receptor (EGFR) dimers, which shows that BS3 can be used as a cross-linking agent for membrane proteins [16]. Additionally, an established protocol for achieving the dimerization of EGFR via cross-linking by BS3has recently been reported [17].

This article reports on oligomeric studies of intra-membrane and extra-membrane Aβ1–42 using BS3 as a cross-linker. Our experimental studies will facilitate further studies of AD pathogenesis.

Materials and methods

Aβ1–42 peptides were purchased from the company (Athens, GA, USA). 1,2-dipalmitoyl-sn-glycero-3-phosphatidylcholine semisynthetic (DPPC, purity> 99%) was purchased from Avanti Polar Lipids (Alabaster, AL). The BS3crosslinker was commercially available from Pierce (Pierce, USA). The mouse monoclonal antibody, 4G8, was purchased from Covance (Covance, USA), and ECL reagent was purchased from Millipore (Millipore, MA); 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP) and dimethylsulfoxide DMSO were purchased from Sigma. F-12 medium was purchased from BioSource (Ham's F-12, BioSource, Australia). Thetransmission electron microscope was a Tecnai G20 (FEI, Netherlands).

Aβ1–42 monomer treatment and oligomer preparation

A total of 222 μl HFIP was added to 1 mg Aβ, and the solution was mixed with a vortex for 5 min and left overnight at 4°C with slow shaking. The solution was then aliquoted into tubes containing 8.6 μl each, evaporated in N2 flow, and stored at -80°C. ADDL preparation was carried out as follows: F12 medium was added to an Aβ tube to give a final concentration of 100 μM, and the sample was incubated at 4°C for 24 hours [18]. Protofibrils preparation was carried out as follows: Aβ1–42 was added to Tris buffer (50 mM Tris, 100 mM NaCl, pH = 7.4) to a final concentration of 200 μM, and the solution was incubated with slow shaking at room temperature for 24 hours [6].

Crosslinking and electrophoresis

First, 5 μM Aβ1–42 and 0.3 mM BS3 or 0.3 mM glutaraldehyde were added to PBS buffer, and the mixtures were allowed to crosslink in ice bath for varying durations. The reaction was then terminated with 1 M Tris (pH7.4). The samples were heated in 5×sample loading buffer at 90°C for 5 minutes before electrophoresis. During electrophoresis, a 4% stacking gel and a 16.5% Tris-tricine separation gel were used.

Liposome preparation

First, 1 ml of phospholipids was vacuum-dried to remove solvents in nitrogen air overnight. Then, 1,000 μl PBS, which was preheated to 70°C, was added to the phospholipids, and the samples were incubated in a 70°C water bath after they were mixed with vortex oscillation. During incubation, the solution was removed from the water bath several times to be vortexed. After incubation, the solution was sonicated for 15 s and subjected to at least 5 freeze-thaw cycles. Liposomes were formed once the solution passed through a filter at temperatures above the phase-transition temperatures of the phospholipids [19].

Transmission electron microscopy (TEM) sample preparation

First, 2 μl Aβ42 solution was added to discharge/hydrophilic-treated carbon films and incubated for 1 min. After washing with ddH2O, the carbon films were negatively stained with 1% sodium phosphotungstate for 1 min. Then, excess liquids were removed, and the carbon films were air-dried for TEM observation.

SEC

SEC was carried out as previously described [20], with some modifications. A Biologic Quadtec UV-Vis Detector (Bio-Rad Laboratories, Hercules, CA, USA) and Superdex 75 10/300 GL column (Tricorn) were used. Columns were eluted at a flow rate of 0.5 mL/min. Aβ40 samples were centrifuged for 10 min at 15,000 × g and 4°C), and the supernatant (50 μL) was then injected into the column. Samples were fractionated at a flow rate of 0.75 mL/min, and peptides were detected by measuring the UV absorbance at 210 nm. Each experiment was performed at least three times.

Results

Principle of the BS3 crosslinking reaction

Di-succinimidylsuberate(DSS) is a water-insoluble N-hydroxysuccinimide ester (NHS-ester), and BS3, shown in Fig 1, is its water-soluble analog.

thumbnail
Fig 1. Molecular structures of BS3.

Molecular structures of di-succinimidylsuberate (disuccinimidyl suberate, DSS) and bis-succinimidylsuberate sodium salt (Bis(sulfosuccinimidyl, BS3).

https://doi.org/10.1371/journal.pone.0173871.g001

Both of them are α-amine-reactive, and although α-amines present on the N-termini of peptides react with NHS-esters, they are seldom available on proteins. Therefore, the reaction with side chains of peptides is the majorreaction that occurs. Furthermore, while five amino acids have nitrogen in their side chains, only the ε-amine of lysine will react to produce a stable product. When BS3 dissolved, its N-containing portion is rapidly removed, resulting in α-amino acids or lysine at the two terminals of this compound. The NHS-Ester Reaction Scheme of amino acids crosslinking with BS3 is described as S1 Fig.

Aβ1–42 has two residues, Lysine-16 and Lysine-28, which can be crosslinked to BS3. The hydrolysis reaction is a competitive reaction and can be terminated by 1 M Tris HCl or 1 M Glycine solution. This NHS-Ester Reaction (S1 Fig) has been previously described in the literature [17,21,22].

In this study, BS3 was adopted to cross-link Aβ1–42. The results are as follows.

Controlling the crosslinking of Aβ monomers by BS3

According to previous reports, membrane-interacting Aβ1–40 can be crosslinked with 12 mM glutaraldehyde for 10 minutes [11]. To avoid excessive crosslinking, we used significantly lower concentration gradients of 0.3 mM, 0.6 mM and 1.2 mM and much shorter time gradients in an ice bath. The Aβ concentration was kept at 5 μM. The result showed that the Aβ1–40 and Aβ1–42 monomer could also readily form several oligomeric forms (Fig 2A, Fig 2B).

thumbnail
Fig 2. Aβ monomers cross-linked with glutaraldehyde and BS3.

(A) Aμ1–42 and (B) Aμ1–40 monomer cross-linked with glutaraldehyde under ice bath conditions. The final concentrations of glutaraldehyde were 0.3 mM, 0.6 mM and 1.2 mM, and 1 minute, 5 minute and 10 minute incubation periods were used. (C) Aμ1–42 and (D) Aμ1–40 monomer cross-linked with BS3 under ice bath conditions. The final concentration of BS3 were 0.3 mM, 0.6 mM and 1.2 mM, and 1 minute, 5 minute and 10 minute incubation periods were used. A 5 μM final concentration of Aβ was used in all the glutaraldehyde and BS3 crosslinking experiments. A total of 1 μg sample was added to each electrophoresis lane, and the 16.5%Tris-Tricine gel was analyzed using silver staining. The concentration of Aβ42 (E) and Aβ40 (F) were the same as the crosslinking concentration. Samples were incubated in an ice bath for 1, 5, or 10 min.

https://doi.org/10.1371/journal.pone.0173871.g002

The ability of glutaraldehyde and BS3 to crosslink the Aβ1–42 monomer was compared for different concentration gradients and time gradients. The results are shown in Fig 2. When the lowest concentration (0.3 mM) and short crosslinking duration (1 minute) was used, glutaraldehyde crosslinked monomers into oligomers, whereas BS3 did not result in excessive crosslinking, even after a crosslinking duration of 5 minute. However, different concentrations of glutaraldehyde and BS3 barely affected crosslinking. It can be seen from Fig 2D that crosslinking was weaker when 1.2 mM of crosslinkers was used, as compared to when 0.6 mM of crosslinkers was used (refer to the discussion section).

BS3 is more sensitive for Aβ oligomeric studies, compared to glutaraldehyde, as a crosslinker

ADDL and protofibril are the two most common Aβ1–42 aggregates [6]. Aβ monomers, ADDL and protofibril were analyzed with a 16.5% Tris-tricine gel. As shown in Fig 3. In the absence of crosslinking, the gray value percentages of aggregate bands were determined in monomers, ADDL, and protofibril. Aggregates of monomers showed reduced percentages by silver staining and western blotting. However, the numbers of Aβ1–42 bands did not differ significantly for the three forms. When 0.3 mM glutaraldehyde was used to crosslink the Aβ1–42 monomer, ADDL, and protofibrils for 1 min, this reagent caused excessive crosslinking overall, and therefore, it was difficult to distinguish the differences among the three Aβ1–42 samples based on the gray values and band numbers. After crosslinking with 0.3 mM BS3 for 5 min, different Aβ1–42 forms, which showed contrasting electrophoresis bands, were present. The monomer and the ADDL sample showed monomer, trimer and tetramer bands when silver staining and western blotting, the protofibril sample showed monomer, trimer, tetramer and higher oligomer bands; for protofibril, the percentage of tetramers and trimers was reduced, while that of the higher oligomers was significantly increased. There were also major differences in the percentages of aggregate gray values between the three samples. Therefore, as a crosslinker, BS3 was much better in distinguishing different oligomers in Aβ samples, as compared to glutaraldehyde. Additionally, even higher oligomers may have been present.

thumbnail
Fig 3. Aβ1–42 monomers, ADDL and protofibrils were cross-linked with glutaraldehyde or BS3.

All experiments were carried out under ice bath conditions. The samples were analyzed on a 16.5% Tris-tricine gel. (A) In the absence of crosslinking, three forms of Aβ1–42. Aβ1–42 was crosslinked with 0.3 mM glutaraldehyde for 1 minute.Aβ1–42 was cross-linked with 0.3 mM BS3 for 5 minutes, and 1 μg sample was added to each electrophoresis lane. The gels were analyzed using silver staining. The "←" indicates the position of the aggregates in the protofibril lane. (B) The same crosslinking reaction conditions as “(A)” were used. But in "(B)", the amount of protein used in each electrophoresis lane was 10 ng. The results were analyzed using the murine monoclonal antibody, 4G8. (C) Histogram showing the grayscale value of each band in (A). “Aggregates” indicates the sum of all other bands excluding monomers. (D) Histogram showing the grayscale value of each band in (B).

https://doi.org/10.1371/journal.pone.0173871.g003

Previous studies have shown that Aβ1–42 oligomerization can be induced through membrane insertion or membrane adsorption. Aggregation mechanisms include membranes as a reaction platform to accelerate Aβ aggregation, electrostatic effects, and Aβ conformational changes, resulting in the formation of ion channels [8,2325]. In order to analyze the state of Aβ1–42 when interacting with a membrane, Aβ1–42 was added to a DPPC liposome solution. A final concentration of 1 μM Aβ1–42 monomer was used, and the mixture was incubated for two hours at room temperature. After centrifugation at 15,000 g × 10 min to remove large aggregates (fibers), the resulting supernatant was analyzed using Transmission electron microscopy (TEM), as shown in Fig 4D and 4E(more seen S3 Fig).The supernatant was subjected to additional ultracentrifugation step at 200,000 g × 30 min. In the samples in which no crosslinking had occurred, the electrophoretic bands for the Aβ1–42 monomers and the precipitants of the Aβ1-42-liposome showed no significant difference, as shown in Fig 4A. For the samples that had been crosslinked with BS3, the precipitant and supernatant of the Aβ1-42-liposome displayed different bands in SDS-PAGE, as indicated in Fig 4B (silver staining) and 4C (Western Blot analysis). Compared to non-crosslinked cases, BS3 crosslinking produced a significant amount of oligomeric Aβ1–42. The same conclusion was reached based on the silver staining and Western Blot analysis. Taken together, these results suggest that as compared to glutaraldehyde, using as a crosslinker BS3 is applicable not only to different oligomeric states of Aβ1–42, but also to Aβ1–42 in the context of a membrane.

thumbnail
Fig 4. BS3 crosslinking to analyze the behavior of Aβ1–42 interaction with liposomes

(A) The Aβ1–42 monomer was incubated with liposomes in the absence of crosslinking. The samples were assessed by electrophoresis and silver stain analysis. There was no significant difference when comparing the Aβ1–42 monomer with "Aβ1–42 monomer + liposome". (B) The Aβ1–42 monomer was added to liposomes, and the mixture was crosslinked with 3 mM BS3 for 5 minutes under ice bath conditions. Compared with the Aβ1–42 monomer, the "Aβ1–42 monomer + liposome" precipitate, but not the supernatant samples, clearly formed oligomers. A total of 1 μg of sample was added to each electrophoresis lane. (C) All experimental conditions were the same as "B", but the amount of protein used in each electrophoresis lane was 10 ng. The results were analyzed by a murine monoclonal antibody, 4G8, and Western blot. (D) Aβ1–42 DPPC was incubated with liposomes for 2 hours. The staff was 200 nm, and the liposomes had a diameter of 300 nm. After discarding the Aβ1–42 fiber, the samples were centrifuged and resuspended. The samples without crosslinking were assessed by TEM. (E) All experimental conditions were the same as "(D)", but the sample was crosslinked with 3 mM BS3 before TEM sample preparation. The staff was 100 nm.

https://doi.org/10.1371/journal.pone.0173871.g004

BS3-Aβ1–42 crosslinking observed by TEM

The Aβ1–42 monomer, ADDL and protofibril were assessed by TEM. The Aβ1–42 monomer was only one nanometer, and it was difficult to distinguish the non-crosslinked form from that which had crosslinked with BS3 (Fig 5A and 5D). Compared to Fig 5E and 5F, Fig 5G and 5H showed considerable Aβ1–42 aggregates after crosslinking with BS3. When glutaraldehyde was used, severe crosslinking occurred. In contrast, BS3 caused only mild crosslinking of Aβ1–42, and these conclusions are similar to those obtained from the electrophoresis results.

thumbnail
Fig 5. Electron microscope photographs of Aβ1–42 monomer, ADDL and protofibril.

(A), (B) and (C) show the non-crosslinked samples. (D), (E) and (F) show the samples that had been crosslinked with 0.3 mM BS3 for 5 minute. (G), (H) and (I) are samples that had been crosslinked with 0.3 mM glutaraldehyde for 1 minute. The staff was 50 nm.

https://doi.org/10.1371/journal.pone.0173871.g005

Analysis of BS3-Aβ crosslinking by SEC

Aβ monomers and aggregates exist simultaneously in nature before and after crosslinking. SDS-PAGE was used to characterize these monomers and aggregates, as described above. However, since many Aβ mixtures are SDS-labile, the use of SDS-PAGE can lead to some problems in the analysis. Accordingly, the use of SEC combined with SDS-PAGE may be appropriate because SEC fractionates Aβ samples under nondenaturing conditions on the basis of molecular weight. SEC has been used to analyze different aggregation states of Aβ in solution [20,26].

Glutaraldehyde- and BS3-crosslinked Aβ samples showed different peak shapes from SECunder nondenaturing conditions. The Aβ40 monomer and protofibrils were treated with 6 M guanidine hydrochloride, DMSO, 0.3 mM glutaraldehyde, or BS3 and analyzed on Superdex 75 10/300 GL columns. The results are shown in Fig 6. Aβ40 monomer samples treated with guanidine hydrochloride had only one monomer peak, whereas the DMSO-treated monomer sample showed two oligomer peaks (one small and one large). In the absence of crosslinking, the Aβ40 oligomer sample showed a monomer peak and an oligomer peak. The Aβ40 agglutination diversity pattern was not separated, consistent with previous reports showing that low-molecular-weight oligomers are in rapid equilibrium with monomers and cannot be distinguished [27,28]. Glutaraldehyde-crosslinked monomer and oligomeric samples showed oligomeric peak area ratios that were larger than those of BS3-crosslinked samples. For each injection, the total (monomer + aggregate) peak area was expressed as 100%. For monomer samples, the glutaraldehyde-crosslinked oligomeric peak area was 27.77%, whereas the BS3-crosslinked oligomeric peak area was 15.25%. For oligomeric samples, the glutaraldehyde-crosslinked oligomeric peak area was 30.37%, whereas the BS3-crosslinked oligomeric peak area was 25.03%. These data suggested that the intensity of glutaraldehyde crosslinking was stronger than that of BS3, consistent with the above electrophoresis and TEM data.

thumbnail
Fig 6. Analysis of BS3-Aβ crosslinking by SEC on a Superdex 75 10/300 GL SEC column.

(A) Aβ40 monomer (red line: DMSO solubilization; dark blue line: 6 M guanidine-HCl solubilization; sapphire line: 0.3 mM BS3 crosslinking; black line: 0.3mM glutaraldehyde crosslinking). Fifty microliters was used for each sample (Aβ40 concentration: 150 μM). (B) Aβ40 protofibrils (red line: no crosslinking; sapphire line: 0.3 mM BS3 crosslinking; pink line: 0.3 mM glutaraldehyde crosslinking). Aβ40 monomer (black line: 6 M guanidine-HCl solubilization). Fifty microliters was used for each sample (Aβ40 concentration: 300 μM). (C) Analysis of the peak area for the elution of monomeric Aβ40 (crosslinked or not crosslinked) from monomer samples. Each sample was injected three times. For each sample analysis, the total (aggregate + monomer) peak area was expressed as 100%. For Aβ40 monomers without crosslinking, the monomer:aggregate ratio was 93%:7%. For Aβ40 monomers with BS3 crosslinking, the monomer:aggregate ratio was 84.87%:15.13%. For Aβ40 monomers with glutaraldehyde crosslinking, the monomer:aggregate ratio was 72%:28%. (D) Analysis of the peak area for the elution of protofibril Aβ40 (crosslinked or not crosslinked) from monomer samples. Each sample was injected three times. For each sample analysis, the total (aggregate + monomer) peak area was expressed as 100%. For Aβ40 protofibrils without crosslinking, the monomer:aggregate ratio was 85.3%:14.7%. For Aβ40 protofibrils with BS3 crosslinking, the monomer:aggregate ratio was 75%:25%. For Aβ40 protofibrils with glutaraldehyde crosslinking, the monomer:aggregate ratio was 69.6%:30.4%.

https://doi.org/10.1371/journal.pone.0173871.g006

Based on the above results, BS3 is very useful for Aβ1–42 electrophoresis and electron microscopy analysis in the following aspects: 1) the moderate crosslinking of Aβ1–42 by BS3 helps identify different oligomers in Aβ monomers, ADDL and other forms aggregates through electrophoresis; 2) Compared to glutaraldehyde as a crosslinker, BS3 is more appropriate, as it does not lead to excessive crosslinking; 3) BS3 can be used to crosslink Aβ in the context of a membrane.

Discussion

Soluble Aβ oligomers play an important role in the etiology of the AD process, and thus, in recent decades, the toxic effects of Aβ oligomers have been the subject of considerable research [29,30]. The study of Aβ often requires the detection of Aβ aggregates in order to distinguish monomeric, oligomeric or protofibrilic protein status, which is technically challenging. This same problem exists in the studies of other proteins prone to aggregation, such as α- synuclein, islet amyloid polypeptide (IAPP), human insulin and prion peptide [30,31]. in this paper, we assessed a method of oligomer detection using two common Aβ aggregates, ADDL and protofibril [7], as examples.

Electrophoresis analysis is one of the most commonly used methods in Aβ aggregate detection. However, the use of a denaturing gel results in altered Aβ oligomers, which no longer reflect the original aggregates. Glutaraldehyde has previously been used as a crosslinking agent, such as in the preparation of a decellularized vascular matrix by co-crosslinking procyanidins and glutaraldehyde [32]. However, when glutaraldehyde is used, the excessive cross-linking of Aβ occurs easily, even at very low concentrations and with short crosslinking times [33].

When glutaraldehyde was used, we observed severe aggregation, even for Aβ monomers (Fig 2 and S2 Fig), thus making it hard to differentiate the Aβ oligomer states. When BS3 was used, the crosslinking was relatively "mild", and excessive crosslinking of the Aβ1–42 monomer could be prevented, as long the process was kept to under 5 minutes, as shown in Figs 2 and 3. According to Fig 2, the crosslinking was weaker when the concentration of BS3 was increased to 1.2 mM and the duration was increased to 10min. This is because the large aggregates that had formed at higher BS3concentrations did not enter the gel during electrophoresis. Additionally, the intermediate portion of the gel was advantageous in silver staining, to a certain extent, since the edge strip dye was easily eluted when the entire acrylamide gel, which was placed in a Petri dish, was incubated with rocking on a rotary shaker during the dyeing-destaining process. The concentration of protein used in this study was 5 μM. If a higher protein concentration is required, a corresponding proportional increase in the concentration of BS3 would be required.In the experiments with liposomes, because the preliminary experiments showed that lipids have a greater tendency for interfering with crosslinking, BS3 was added to a final concentration of 3 mM in phospholipid membranes which was ten times of the concentration of BS3 in solution, as shown in Fig 4. Preliminary experiments using different concentration proportion of Aβ1–42 and phospholipid did not yield significantly different results.

Treatment of Aβ1–42 with glutaraldehyde appeared to induce the formation of trimers and tetramers, whereas treatment of Aβ1–42 with BS3 induced the formation of dimers and trimers. This may be due to the generation of different forms of aggregates induced by excessive crosslinking of glutaraldehyde; these aggregates may become more stable trimers and tetramers by extrusion through the electrophoresis gel pores. This concept of gel-induced oligomerization has been reported in a previous work [8] and can be explained by the ability of glutaraldehyde to produce the acetal reaction easily following self-aggregation, resulting in trimers and tetramers. BS3 is the most likely to remove the N group at both ends, thereby producing two dimers through the action of two Aβ molecules. One of the two dimers forms through Lys and then interacts with a BS3 molecule to form a trimer.

Our results showed that the trimers and tetramers in ADDL and protofibril samples that had not been crosslinked were more abundant than those in the monomer sample. This result could be explained by the observation that the larger aggregates may become more stable trimers and tetramers by extrusion through electrophoresis gel pores [8]. When BS3 was used as a crosslinker, few tetramers were found in the protofibril samples because BS3 stabilized the aggregates without inducing the formation of tetramers.

Gel electrophoresis was performed under denaturing conditions. However, SEC analyzes the crosslinking of proteins under nondenaturing conditions. Our results also showed that glutaraldehyde had a higher crosslinking capacity than BS3. Notably, in the SEC experiment, we did not use Aβ1–42 because Aβ1–42 is aggregates easily and cannot be easily maintained in the monomer state. Additionally, in UA210, the light absorption value is low, resulting in prevention of protein waste. This further supported the use of Aβ1–40 in this study. We did not use Aβ-40 ADDL because ADDL is prepared in culture medium, which contains many light-absorbing groups(S4 Fig). Therefore, in this study, we used Aβ1–40.

In short, for Aβ, which are prone to aggregate, electrophoresis after BS3 crosslinking enables the detection of discrete protein bands that correspond to different aggregates of Aβ. This suggests that BS3 crosslinking suitable for oligomer detection of Aβ from various sources, including chemically synthesized, naturally expressed and liposome-interacting Aβ. Therefore, BS3 crosslinking should be useful in the study of the toxicity mechanism of Aβ and AD.

Supporting information

S1 Fig. The NHS-Ester Reaction Scheme of amino acids crosslinking with BS3.

https://doi.org/10.1371/journal.pone.0173871.s001

(TIF)

S2 Fig. Aβ1–42 monomers, ADDL and protofibrils were cross-linked with glutaraldehyde or BS3.

https://doi.org/10.1371/journal.pone.0173871.s002

(TIF)

S3 Fig. Different Aβ1–42 forms and liposome samples were observed by TEM.

https://doi.org/10.1371/journal.pone.0173871.s003

(TIF)

S4 Fig. Analysis of Aβ1–42 monomer, ADDL, and F12 culture medium by SEC.

https://doi.org/10.1371/journal.pone.0173871.s004

(TIF)

Author Contributions

  1. Conceptualization: JMS.
  2. Data curation: LZ.
  3. Formal analysis: JMS.
  4. Funding acquisition: EQL JMS.
  5. Investigation: LZ JMS JP.
  6. Methodology: JMS.
  7. Project administration: EQL.
  8. Resources: EQL.
  9. Software: JMS.
  10. Supervision: LZ.
  11. Validation: JP LZ.
  12. Visualization: LZ.
  13. Writing – original draft: JMS.
  14. Writing – review & editing: EQL.

References

  1. 1. Prince M, Bryce R, Albanese E, Wimo A, Ribeiro W, Ferri CP. The global prevalence of dementia: a systematic review and metaanalysis. Alzheimers Dement. 2013;9: 63–75. pmid:23305823
  2. 2. Reitz C, Brayne C, Mayeux R. Epidemiology of Alzheimer disease. Nat Rev Neurol. 2011;7: 137–152. pmid:21304480
  3. 3. Tarasoff-Conway JM, Carare RO, Osorio RS, Glodzik L, Butler T, Fieremans E, et al. Clearance systems in the brain-implications for Alzheimer disease. Nat Rev Neurol. 2015;11: 457–470. pmid:26195256
  4. 4. Fortea J, Vilaplana E, Alcolea D, Carmona-Iragui M, Sanchez-Saudinos MB, Sala I, et al. Cerebrospinal fluid β-amyloid and phospho-tau biomarker interactions affecting brain structure in preclinical Alzheimer disease. Ann Neurol. 2014;76: 223–230. pmid:24852682
  5. 5. Melah KE, Lu SY, Hoscheidt SM, Alexander AL, Adluru N, Destiche DJ, et al. Cerebrospinal fluid markers of Alzheimer's disease pathology and microglial activation are associated with altered white matter microstructure in asymptomatic adults at risk for Alzheimer's disease. J Alzheimers Dis. 2016;50: 873–886. pmid:26836182
  6. 6. Ahmed M, Davis J, Aucoin D, Sato T, Ahuja S, Aimoto S, et al. Structural conversion of neurotoxic amyloid-beta(1–42) oligomers to fibrils. Nat Struct Mol Biol. 2010;17: 561–567. pmid:20383142
  7. 7. Benilova I, Karran E, De Strooper B. The toxic Aβ oligomer and Alzheimer's disease: an emperor in need of clothes. Nat Neurosci. 2012;15: 349–357. pmid:22286176
  8. 8. Zhang YJ, Shi JM, Bai CJ, Wang H, Li HY, Wu Y, et al. Intra-membrane oligomerization and extra-membrane oligomerization of amyloid-beta peptide are competing processes as a result of distinct patterns of motif interplay. J Biol Chem. 2012;287: 748–756. pmid:22105077
  9. 9. Jan A, Gokce O, Luthi-Carter R, Lashuel HA. The ratio of monomeric to aggregated forms of Abeta40 and Abeta42 is an important determinant of amyloid-beta aggregation, fibrillogenesis, and toxicity. J Biol Chem. 2008;283: 28176–28189. pmid:18694930
  10. 10. Klein WL. Synaptotoxic amyloid-β oligomers: a molecular basis for the cause, diagnosis, and treatment of Alzheimer's disease? J Alzheimers Dis. 2013;33: S49–65. pmid:22785404
  11. 11. Quist A, Doudevski I, Lin H, Azimova R, Ng D, Frangione B, et al. Amyloid ion channels: a common structural link for protein-misfolding disease. Proc Natl Acad Sci U S A 2005;102: 10427–10432. pmid:16020533
  12. 12. Bitan G, Kirkitadze MD, Lomakin A, Vollers SS, Benedek GB, et al. Amyloid beta -protein (Abeta) assembly: Abeta 40 and Abeta 42 oligomerize through distinct pathways. Proc Natl Acad Sci U S A 2003;100: 330–335. pmid:12506200
  13. 13. Levine H 3rd. Soluble multimeric Alzheimer beta(1–40) pre-amyloid complexes in dilute solution. Neurobiol Aging. 1995;16: 755–764. pmid:8532108
  14. 14. Schmidt C, Robinson CV. A comparative cross-linking strategy to probe conformational changes in protein complexes. Nat Protoc 2014;9: 2224–2236. pmid:25144272
  15. 15. Ivanov KI, Bašić M, Varjosalo M, Mäkinen K. One-step purification of twin-strep-tagged proteins and their complexes on strep-tactin resin cross-linked with bis(sulfosuccinimidyl) suberate (BS3). J Vis Exp. 2014;86: e51536.
  16. 16. Sorkin A, Carpenter G. Dimerization of internalized epidermal growth factor receptors. J Biol Chem. 1991;266: 23453–23460. pmid:1683870
  17. 17. Turk HF, Chapkin RS. Analysis of epidermal growth factor receptor dimerization by BS(3) cross-linking. Methods Mol Biol. 2015;1233: 25–34. pmid:25319886
  18. 18. Lambert MP, Barlow AK, Chromy BA, Edwards C, Freed R, Liosatos M, et al. Diffusible, nonfibrillar ligands derived from Abeta1-42 are potent central nervous system neurotoxins. ProcNatlAcadSci U S A 1998;95: 6448–6453.
  19. 19. Ji SR, Ma L, Bai CJ, Shi JM, Li HY, Potempa LA, et al. Monomeric C-reactive protein activates endothelial cells via interaction with lipid raft microdomains. FASEB J 2009;23: 1806–1816. pmid:19136614
  20. 20. Jan A, Hartley DM, Lashuel HA. Preparation and characterization of toxic Abeta aggregates for structural and functional studies in Alzheimer's disease research. Nat Protoc. 2010;5: 1186–1209. pmid:20539293
  21. 21. Staros JV, Wright RW, Swingle DM. Enhancement by N-hydroxysulfosuccinimide of water-soluble carbodiimide-mediated coupling reactions. Anal Biochem. 1986;156: 220–222. pmid:3740412
  22. 22. Partis MD, Griffiths DG, Roberts GC, Beechey RB. Cross-linking of protein by ω-maleimidoalkanoylN-hydroxysuccinimido esters. J Protein Chem. 1983;2: 263–277.
  23. 23. Murray IV, Liu L, Komatsu H, Uryu K, Xiao G, Lawson JA, et al. Membrane-mediated amyloidogenesis and the promotion of oxidative lipid damage by amyloid beta proteins. J Biol Chem. 2007;282: 9335–9345. pmid:17255094
  24. 24. Bokvist M, Lindstrom F, Watts A, Grobner G. Two types of Alzheimer's beta-amyloid (1–40) peptide membrane interactions: aggregation preventing transmembrane anchoring versus accelerated surface fibril formation. J Mol Biol. 2004;335: 1039–1049. pmid:14698298
  25. 25. Lin H, Bhatia R, Lal R. Amyloid beta protein forms ion channels: implications for Alzheimer's disease pathophysiology. FASEB J 2001;15: 2433–2444. pmid:11689468
  26. 26. Walsh DM, Lomakin A, Benedek GB, Condron MM, Teplow DB. Amyloid beta-protein fibrillogenesis. Detection of a protofibrillar intermediate. J Biol Chem. 1997;272: 22364–22372. pmid:9268388
  27. 27. Bitan G, Teplow DB. Preparation of aggregate-free, low molecular weight amyloid-beta for assembly and toxicity assays. Methods Mol Biol. 2005;299: 3–9. pmid:15980591
  28. 28. Bitan G, Fradinger EA, Spring SM, Teplow DB. Neurotoxic protein oligomers—what you see is not always what you get. Amyloid 2005;12: 88–95. pmid:16011984
  29. 29. Ochiishi T, Doi M, Yamasaki K, Hirose K, Kitamura A, Urabe T, et al. Development of new fusion proteins for visualizing amyloid-β oligomers in vivo. Sci Rep. 2016;6: 22712. pmid:26982553
  30. 30. Jucker M, Walker LC. Self-propagation of pathogenic protein aggregates in neurodegenerative diseases. Nature. 2013;501: 45–51. pmid:24005412
  31. 31. Jucker M, Walker LC. Pathogenic protein seeding in Alzheimer disease and other neurodegenerative disorders. Ann Neurol. 2011;70: 532–540. pmid:22028219
  32. 32. Wang X, Ma B, Chang J. Preparation of decellularized vascular matrix by co-crosslinking of procyanidins and glutaraldehyde. Biomed Mater Eng. 2015;26: 19–30. pmid:26484552
  33. 33. Cheung DT, Nimni ME. Mechanism of crosslinking of proteins by glutaraldehyde II. Reaction with monomeric and polymeric collagen. Connect Tissue Res. 1982;10: 201–216. pmid:6299648