• No results found

Polymorphic assembly of virus-capsid proteins around DNA and the cellular uptake of the resulting particles

N/A
N/A
Protected

Academic year: 2021

Share "Polymorphic assembly of virus-capsid proteins around DNA and the cellular uptake of the resulting particles"

Copied!
13
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Contents lists available atScienceDirect

Journal of Controlled Release

journal homepage:www.elsevier.com/locate/jconrel

Polymorphic assembly of virus-capsid proteins around DNA and the cellular

uptake of the resulting particles

M.V. de Ruiter, R.M. van der Hee, A.J.M. Driessen, E.D. Keurhorst, M. Hamid,

J.J.L.M. Cornelissen

Laboratory for Biomolecular Nanotechnology, MESA+ Institute for Nanotechnology, University of Twente, 7500 AE Enschede, the Netherlands

A R T I C L E I N F O

Keywords: Virus-like particles Transfection Plant virus

Cargo controlled assembly

A B S T R A C T

Virus-like particles (VLPs), i.e. molecular assemblies that resemble the geometry and organization of viruses, are promising platforms for therapeutics and imaging. Understanding the assembly and cellular uptake pathways of VLPs can contribute to the development of new antiviral drugs and new virus-based materials for the delivery of drugs or nucleic acid-based therapies. Here we report the assembly of capsid proteins of the cowpea chlorotic mottle virus (CCMV) around DNA into defined structures at neutral pH. Depending on the type of DNA used, we are able to create spherical structures of various diameters and rods of various lengths. In order to determine the shape dependency, the cellular uptake routes and intracellular positioning of these formed polymorphic VLPs in RAW264.7, HeLa and HEK 293 cells are evaluated usingflow cytometry analysis with specific chemical in-hibitors for different uptake routes. We observed particular uptake routes for the various CCMV-based nanos-tructures, but the experiments point to clathrin-mediated endocytosis as the major route for cell entry for the studied VLPs. Confocal microscopy reveals that the formed VLPs enter the cells, with clear colocalization in the endosomes. The obtained results provide insight in the cargo dependent VLP morphology and increase the understanding of shape dependent uptake into cells, which is relevant in the design of new virus-based structures with applications in drug and gene delivery.

1. Introduction

The uptake of viruses is an essential part in the virus replication life cycle [1]. Evaluation of this viral uptake, its uptake route and their positioning inside cells is important, because increased understanding can help in the development of new treatments for viral diseases [2]. Furthermore, the improved knowledge of virus uptake can be used in the creation and design of optimal virus-based nanostructures for medical applications. After all, the main purpose of a viral protein shell is to act as a delivery vehicle for their nucleic acid genome. They can therefore be exploited to deliver non-native cargos, like drugs, genes and enzymes to cells, for the treatment of various diseases [3–6].

In general, each virus and its cellular uptake is evaluated separately, because of the unique geometrical structure, exterior (epitopes), host cell and uptake route [7]. Because of the structural robustness of most viruses it is challenging to study‘nature's rules’ on the shape dependent uptake of virus-based (or other) nanoparticles [8]. Until now this is mainly evaluated using quantum dots, gold nanoparticles and polymers [9,10]. The results of these examples show that the internalization

pathway and intracellular trafficking is influenced by size [11], surface charge [12], shape [13], surface properties [14], rigidity [15] and composition of nanoparticles [16]. For direct comparison it is important that all of these properties are studied as constant variables.

However, to the best of our knowledge, only one virus has been used to evaluate the shape dependence of viral uptake: the tobacco mosaic virus (TMV). Different morphologies of this virus are evaluated in vitro to determine their uptake mechanism and in vivo to determine their biodistribution [17,18] These studies show that the uptake and uptake pathway of the virus structures is clearly influenced by the aspect ratio used.

In this work we aim to study the shape dependent viral uptake in vitro to confirm the results from TMV with a different virus. We use the cowpea chlorotic mottle virus (CCMV), which is a well-studied model virus to study viral assembly, because it shows controllable disassembly and reassembly behavior. In its native form the protein capsid consists of 90 capsid protein (CP) dimers. Next to this, it is a simple icosahedral virus with minimal safety issues [19–21], which is an advantage when applying such a virus for future medical applications.

https://doi.org/10.1016/j.jconrel.2019.06.019

Received 13 March 2019; Received in revised form 11 June 2019

Corresponding author.

E-mail address:j.j.l.m.cornelissen@utwente.nl(J.J.L.M. Cornelissen).

Journal of Controlled Release 307 (2019) 342–354

Available online 19 June 2019

0168-3659/ © 2019 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/BY-NC-ND/4.0/).

(2)

The native RNA of the CCMV in the relatively large interior cavity can easily be replaced by various other cargos that can eventually be delivered to cells, including inorganic particles [22,23], polymers [24,25] micelles, dyes, [28, 29] drugs, [30,5], nucleic acids [2,31,32] and even active enzymes. [33–36] Such packing can result in poly-morphic assembly of the formed CCMV based virus-like particle (VLPs) which is dependent on the cargo characteristics. For example, when packaging non-native RNA into its capsid this results in spherical VLPs, ranging from Ø = 23 nm particles with an interior of about Ø = 12 nm for 120 nucleotide (nt) cargo, to the native Ø = 28 nm with an interior of about Ø = 18 nm for a 3000 nt strand. When the RNA is longer than 6400 nt, they start forming dimers, trimers and higher ordered clusters of the Ø = 28 nm capsids [37]. In another example CPs can be used to encapsulate DNA-origami structures, clearly showing the different shapes that can be encapsulated by this virus protein [38]. The loading capacity of CCMV-based VLPs is dependent on the shape and size, but pH stable, so-called, T = 1 protein cages with a volume of V ± 268 nm3, have been reported with potentially therapeutic dyes, enzymes and other cargo [22–30,34,36].

We aimed to create CCMV-based VLPs of different size and shape by mixing free CCMV-CPs with different lengths of single stranded DNA (ssDNA) in an attempt to create spherical or icosahedral VLPs and different lengths of double stranded DNA (dsDNA) to create rod-like VLPs (Fig. 1). The higher persistence length of the dsDNA does lead to long and polydisperse rods, which are independent of the length of the cargo [32]. In order to better correlate the length of the rod-like VLPs to the length of the dsDNA, we optimized the assembly conditions by fine-tuning the CP-CP interaction. The addition of MgCl2is an important

parameter in this, as it stabilizes CCMV by complexing with the acidic residues of Glu81 [39].

After the creation of the different CCMV-based nanostructures we evaluated their uptake routes into various cell lines, by using inhibitors for (specific) pathways. We furthermore confirmed the uptake by con-focal microscopy and evaluated the intracellular positioning.

The resulting structures and results can be used for the encapsula-tion of various materials. This may be of great interest for the delivery of antisense oligo DNA, mRNA, gene and small interfering RNA (siRNA) or DNA (siDNA) [6]. Additionally, it can be used to improve en-capsulation of other foreign cargo into viral capsids to develop new materials with applications in, for example, medicine and nano-technology.

2. Materials and methods 2.1. Materials

Chemicals were purchased from Sigma-Aldrich (St Louis, MO, USA) unless stated otherwise and were used without further purification. The single stranded DNA and forward and reverse primers for Taq poly-merase PCR of the different double stranded DNA lengths, were syn-thesized by Eurofins genomics (Ebersberg, Germany). Buffers were prepared using double deionized water from a Millipore Adv. A10 system (18 MΩ cm at 25 °C). The production, purification and char-acterization of the CCMV, the removal of its RNA and the isolation of the CP were carried out according to literature procedures [35]. Only CPs with an absorbance ratio ofλ = 280/260 nm > 1.6 were used for experiments. All reactions were carried out at room temperature and proteins and assemblies were stored at 4 °C.

2.2. Preparation offluorescent CCMV

ATTO 647 N (ATTO-TEC, Siegen, Germany) or Alexafluor 647 (ThermoFisher Scientific, Waltham, USA) NHS ester was dissolved in dry DMSO and mixed in a 0.003 (w/w) ratio with CCMV in 10 mM Phosphate Buffered Saline (PBS) pH 7.4. The unreacted dye was re-moved by dialysis with PBS with 3× refreshment. The degree of la-beling (DOL) was determined by theλ = 260 and 646 nm absorption of the CCMV and dye respectively.ε646 Atto dye= 150,000 cm−1M−1, with

a correction factor of 0.04 at λ = 260 nm (CF260 Atto dye) and ε260 CCMV= 5.87 mL/mg*cm were used as input for the Lambert-Beer

for-mula. ε280 CP= 24,075 cm−1M−1, ε647 dye= 150,000 cm−1M−1.

CF260 dye= 0.03 CF280 dye= 0.03 and CF260 CP= 0.62. The weight of

CP is 20,313 Da and is about 75% of the total mass of the virus. Thus, the degree of labelling is around 0.05 dyes/CP. This is optimized to have maximum fluorescence in the experiments, minimal hindrance during assembly of the viral proteins and to reduce the influence of the dye on the uptake. Upon addition of this dye no aggregation was ob-served. The same batch of CP-dyes were used in the fabrication of the CCMV-DNA based nanostructures, to ensure the same degree of la-beling.

2.3. Polymerase chain reaction

PCR was performed on a primus 25 PCR (Peqlab). Standard condi-tions for taqpolymerase (New England Biolabs, Ipwich, MA, USA) were

Fig. 1. Schematic overview of the used approach to create different viral nanostructures (not to scale). After removal of the native RNA the disassembled CPs are mixed with different lengths of ssDNA or dsDNA to create spherical or rod-like virus-like particles. The 3D structure of the CCMV-based rod-like structures is not confirmed, so an artistic impression is shown.

(3)

used consisting of standard taqpolymerase buffer, 500 μM dNTPs (New England Biolabs), 0.2μM primers and 100 pg pET15-b plasmid DNA as a template (Novagen, Merck, Kenilworth, NJ, USA). We performed 40 cycles that switched between 95 °C, 50 °C and 68 °C, at an interval of

30 s each. The DNA was purified by exchanging the solution 5 times against MilliQ using Amicon Ultra centrifugalfilters (3 K MWCO) and using precipitation in ice-cold ethanol. Analysis with a ThermoFisher Scientific NanoDrop 1000 Spectrophotometer gave the purity and

Fig. 2. Agarose gel electrophoresis on a 1% gel for the analysis of various lengths of ssDNAs mixed with or without CP after 1 week of assembly in assembly buffer. Bands on the gel werefirst visualized with (A) SYBR safe DNA stain, which is followed by visualization with (B) coomassie protein stain. Top lanes from left to right showing (1) 50 bp DNA ladder, the dsDNAs of a length of: (2) 50 bp, (3) 50 bp + CP, (4) 100 bp, (5) 100 bp + CP, (6) 150 bp, (7) 150 bp + CP, (8) 200 bp, (9) 200 bp + CP, (10) 300 bp, (11) 300 bp + CP, (12) 500 bp, (13) 500 bp + CP and (14) CP without any DNA. Bottom lanes from left to right showing (1) 50 bp DNA ladder, the dsDNAs of a length of: (2) 1000 bp, (3) 1000 bp + CP, the ssDNA of (4) 62 nt, (5) 62 nt + CP, (6) sonicated salmon ssDNA (~700 nt), (7) sonicated salmon ssDNA (~700 nt) + CP, (8) calf thymus dsDNA, (9) calf thymus dsDNA + CP, (10) M13MP18 circular plasmid ssDNA (6407 nt), (11) M13MP18 circular plasmid ssDNA (6407 nt) + CP, (12) CP without any DNA, (13) no sample and (14) native CCMV. The red arrow shows the retention of the CP protein band in the wells. (For interpretation of the references to colour in thisfigure legend, the reader is referred to the web version of this article.)

(4)

concentration. The length of the products was verified using agarose gel electrophoresis (AGE).

2.4. Encapsulation of different lengths of ssDNA

CCMV CP in 5× assembly buffer (pH 7.2, 250 mM Tris-HCl buffer containing 250 mM NaCl, 50 mM KCl and 25 mM MgCl2) was diluted to

a concentration of 7.5 mg/mL. The ssDNA of varying lengths of 60 nt, sonicated salmon ssDNA (average 700 nt from Sigma Aldrich) and M13MP18 ssDNA plasmind (New England Biolabs) were dissolved in Milli-Q water to a concentration of 1 mg/mL. For the encapsulation, CCMV-CP and ssDNA were mixed in a 6:1 mass ratio (w/w) of CP to ssDNA and a 4:1 (v/v) ratio of Milli-Q to buffer by adding more Milli-Q, yielding a 1× assembly buffer solution containing 50 mM Tris-HCl, 50 mM NaCl, 10 mM KCl and 5 mM MgCl2with a pH of 7.2. This was

subsequently incubated at least overnight for ssDNA at 4 °C before analysis and purification.

2.5. Encapsulation of different lengths of dsDNA

Double stranded DNA made from PCR with lengths of 50, 100, 200, 500 and 1000 bp was encapsulated using the same method as the ssDNA samples, only having a DNA to CP mass ratio (w/w) of 1:4. For this encapsulation procedure various buffer compositions were tested to optimize the rod formation process. These included pH 7.2 buffers based on Phosphate, 2-(N-morpholino)ethanesulfonic acid (MES), Tris (hydroxymehyl)aminomethane (TRIS) and citric acid buffers, con-taining various concentrations of salt and MgCl2. The earlier mentioned

5× assembly buffer was the most optimal for the assembly. 2.6. Agarose Gel electrophoresis (AGE)

A 1% agarose gel in TAE buffer with either 1× SYBR safe or SYBR Gold (ThermoFisher Scientific) was loaded with 5 μL unpurified sample mixed with 2μL gel loading dye and run for 1 h at 100 V. The gel was

Fig. 3. TEM images of the CCMV assemblies around different nucleotide templates. (A-E) spherical particles formed by scaffolds: (A) native CCMV, (B) 62 nt + CP, (C) salmon ssDNA (700 nt) + CP and (D) 50 bp dsDNA + CP. (E) Histograms of the size distribution of spherical particles based on the TEM data. (F-J) The rod-like assemblies in the presence of rods assembled in the presence of Mg2+. (F) 100 bp + CP, (G) 200 bp + CP, (H) 500 bp + CP and (I) 1000 bp + CP, (J) Size dis-tribution of rod-like assemblies based on TEM data.

(5)

imaged using a Gel Doc™ EZ on a UV tray (Bio-Rad, Hercules, CA, USA). The gel was then shortly destained in Milli-Q, followed by 12 h in-cubation with Bio-Safe Coomassie G-250 stain (Bio-Rad). Excess dye was removed by destaining overnight in Milli-Q. It was again imaged on a Gel Doc™ EZ, this time on a wide field tray, for determination of protein content.

2.7. Size-exclusion chromatography (SEC)

Analysis and purification of the spherical VLPs was performed by SEC using a Superose 6 10/300 GL column (GE Healthcare, Chicago, IL, USA) on a FPLC-system (Acta Purifier, GE Healthcare), eluting at 0.5 mL/min with 1× assembly buffer. Absorption was monitored at λ = 260 nm and 280 nm.

2.8. Dynamic light scattering (DLS)

The size distribution of VLPs in assembly buffer pH 7.2 was mea-sured with dynamic light scattering (DLS) using a Microtrac Nanotrac Wave W3043 (Krefeld, Germany). Viscosity and refractive index of water and the refractive index of native CCMV (1.54) were used.

2.9. Transmission electron microscopy (TEM)

Samples (5μL) were applied onto Formvar-carbon coated grids (FCF-200 mesh copper grid, Electron Microscopy Sciences). After 1 min, the excess of liquid was drained usingfilter paper. Uranyl acetate (5 μL, 1% w/v) was added and the excess of liquid was drained after 15 s and dried for 30 min at room temperature. The samples were examined on a FEG-TEM (Phillips CM 30) operated at 300 kV acceleration voltages. The size distribution was measured with ImageJ. To test the influence of the medium on the virus and VLPs, we imaged them with TEM after 4 h incubation in DMEM at a concentration of 1 mg/mL (Fig. S6).

2.10. SDS-PAGE

10μL of sample was mixed with 9 μL of sample buffer and 1 μL 2-mercaptoethanol. The mixture was heated at 99 °C for 5 min to de-nature the protein, after which the mixture was used tofill the wells of 4–15% stain free precast polyacrylamide gels (Bio-Rad). Precision Plus Protein™ Unstained Protein Standard was added in a separate well. Electrophoreses was conducted at 100 V for 5 min followed by 200 V for approximately 20 min. Gels where activated with UV for 5 min on a stain-free enabled UV transilluminator and imaged with a Gel Doc™ EZ system with Image Lab software (Bio-Rad). Using 5 different con-centrations of CP as reference, the CP content of the purified samples was determined.

2.11. Cell culture

HeLa, RAW264.7 and HEK293 cells were grown in Dulbecco's modified Eagle's medium (DMEM) + 10% fetal bovine serum (FBS) + penicillin 100 U/mL and streptomycin 100μg/mL at 37 °C and 5% CO2in T75flasks (Corning). The cells were passed using trypsin,

versene and PBS respectively and counted using a Bürker-Türk counting chamber.

2.12. Flow cytometry with uptake inhibition

During theflow cytometry measurements to determine uptake ef-ficiency and uptake pathway, the cells were seeded at 5 * 104cells/well

in a 48 well plate and left to adhere overnight. The next morning, the cells used for the uptake pathway experiments were incubated with specific inhibitors as shown in Table S1 which are in the range of concentrations that are used in earlier research. [17,40] After ½ hour of incubation with the inhibitors, the positive controls or CCMV nanos-tructures were added. Inhibitor activity was tested and optimized using fluorescently labeled Transferrin From Human Serum Alexa Fluor™ 647 Conjugate (Fisherscientific), recombinant Cholera toxin Subunit B

Fig. 4. Schematic representation of some of the different uptake routes into cells. Showing the chemical inhibitors, the pathway(s) they block and the positive controls used.

(6)

Alexa Fluor™ 647 Conjugate (Fisher Scientific), Dextran Alexa Fluor™ 647; 10,000 MW (Fisher Scientific) and Latex beads, carboxylate and fluorescein-modified polystyrene (Sigma-Aldrich). To evaluate the up-take of the different virus-based nanostructures, 16 μg of the different viral constructs was added directly to the medium. Which corresponds to approximately 4*107of the native CCMV particles per cell. After 4 h

incubation, the cells were washed with PBS, removed from the plate with trypsin, centrifuged (14,000 g, 3 min) and resuspended in life cell imaging solution (HEPES buffer from Fisher Scientific) and kept on ice. Viability staining (CellTrace™ Calcein Violet AM, Fisher Scientific) was added 20 min before analyzing and dead staining (SYTOX™ Green Nu-cleic Acid Stain - 5 mM Solution in DMSO, Fisher Scientific) was added 5 min before analyzing according to the supplier's protocols. The cells were analyzed in triplo, in 3 separate experiments on a BD FACS ARIA II using a λ = 375 nm laser with a 450/40 filter for the calcein violet, λ = 488 nm with 530/30 filter for the fluorescein product of the sub-strate, 488 nm with 616/23filter for the PI and using the λ = 633 nm laser with 660/20filter for the ATTO 647 N labeled CCMV. Data was processed, to discriminate between living, viable and dead cells using flowing software to extract the average relative fluorescent intensity of the living and viable cells for each of the used conditions. This intensity of the sample was normalized by dividing the intensity of viral nanos-tructures with inhibitors added by the intensity of the nanosnanos-tructures

without inhibitor added. The bar graphs were plotted in Origin. 2.13. Confocal microscopy

For imaging with a confocal microscope, the cells were seeded in medium at 5000 cells/well in a 96-well sensoplate (Greiner BioOne, Kremsmunster, Austria). To each well 1μL of Bacman 2.0 (ThermoFisher Scientific) was added, which encode for fluorescent protein organelle stains for: Golgi (GFP), Mitochondria (GFP), Lysosomes (RFP), early endosome (RFP) and late endosome (RFP). Cells were incubated for 4 h, washed and subsequently incubated with 1000× diluted Bacman Enhancer (ThermoFisher Scientific) for 60 min in medium. The cells were washed again and left overnight with fresh medium at 37 °C, 5% CO2. The following day, 4μg of the virus like

particle constructs was added to each well (equivalent to approximately 1*107native CCMV particles per cell) and incubated for 4 h follwed by

washing twice with Hanks balanced salt solution (HBSS from Fisher Scientific). The cells were fixed for 10 min with 4% paraformaldehyde at room temperature and washed twice with HBSS. The nucleus was stained with Dapi (50 ng) for 5 min and washed with PBS. The mem-brane was stained with 250 ng CF405S Wheat Germ Agglutinin (Biotium) for 20 min and washed with PBS. The plates were stored in the dark at 4 °C in PBS. The cells were measured independently at 4

Fig. 5. Uptake of the native CCMV with an ATTO 647 Nfluorescent dye after 4 h incubation in different cells lines. The uptake and standard deviation is determined using the averagefluorescent intensity of the life cells as measured by flow cytometry, with n = 9. (A) Absolute comparison of CCMV uptake for each cell-line. (B-D) show the uptake after pre-incubation with various uptake inhibitors. The uptake was normalized with CCMV without inhibitor set at 100%. (B) RAW 264.7 macrophages, (C) HeLa cancer cells and (D) HEK 293 kidney cells. *significant inhibition compared with uninhibited CCMV (P < .05).

(7)

(caption on next page)

(8)

separate channels using an OPERA automated high content, spinning disk confocal imager (PerkinElmer, Waltman, MA, USA) equipped with λ = 405, 488, 565 and 633 nm lasers, which was pre-calibrated using a calibration plate withfluorescent beads from PerkinElmer. The images were scaled and merged using Matlab. The Pearson coefficients were determined with ImageJ, using cell-by-cell intensity plot profiles of the separate colored images, from which the Pearson coefficient was cal-culated from at least 10 different profiles.

3. Results and discussion

3.1. DNA to create different nanostructures with CCMV-CP

Different lengths of purified dsDNA from polymerase chain reaction (PCR), commercial sonicated ssDNA from salmon (average 700 nt), 62 nt ssDNA (Eurofins), M13MP18 ssDNA plasmid (circular, 6407 nt) and dsDNA from calf thymus were mixed separately with the free CP dimers in assembly buffer (see the Experimental Section for details). A DNA to CP mass ratio of 1 to 6 was used for the encapsulation of the ssDNA. This ‘golden ratio’ was also reported in literature for the en-capsulation of RNA, [37] while for dsDNA a ratio of 1 to 4 was used during the encapsulation. The buffer composition and ratios were chosen such to have the highest yield of particles and the best-defined rods. After > 16 h incubation at 4 °C the assemblies were analyzed on a 1% agarose gel (Fig. 2).

The agarose gel (Fig. 2A) shows that when CP is added to the DNA, the materials migrates substantially less far into the gel compared to the samples without CP. The lower mobility of the mixtures in the gel in-dicates a morphology change, which is expected upon VLP formation. This is especially clear for the mixtures containing dsDNA, which mi-grates only slightly into the gel. Similarly, the difference between the migration of the CP and the CCMV virion is caused by the presence of negative charge resulting from the encapsulated RNA in the latter. In-terestingly, the ssDNA samples (62 nt, salmon ssDNA and the M13MP18 ssDNA plasmid) had only a minor mobility shift and could penetrate much further into the gel. Also, the native virus could migrate further into the gel. All these ssDNA samples were expected to form spherical VLPs. Apparently, these supposed spherical structures show different migration behavior compared to the samples that have an expected rod-like shape, i.e. that are composed of CP and dsDNA. When comparing Fig. 2A withFig. 2B it can be seen that there is colocalization of pro-teins with the DNA, indicating that the decrease in electrophoretic mobility is caused by the formation of DNA-CP complexes.

Next, the various assemblies were analyzed using transmission electron microscopy (TEM,Fig. 3). As expected, nicely formed spherical VLPs could be observed for the samples of ssDNA mixed with CP. The average exterior diameters of the particles where Ø = 28 ± 1 nm, 24 ± 3 nm and 21 ± 2 nm for the native (RNAfilled) CCMV, salmon ssDNA (700 nt average) and 62 nt ssDNA, respectively. Interestingly, short 50 bp dsDNA mixed with CP, also resulted in spherical structures with Ø = 25 nm ± 3 nm. This is likely caused by the short length of the DNA, which is about 17 nm in a fully stretched conformation. The interior diameter of the native T = 3 capsid is close to 18 nm, so the DNA can‘fit’ inside the native capsid and in that way induce the for-mation of spherical structures. The particles with the ssDNA scaffolds were purified using size exclusion chromatography (SEC) and

additionally characterized in solution with dynamic light scattering (DLS) (Fig. S1). These results confirmed the formation of structures with an average diameter of Ø =27 ± 1 nm, 25 ± 2 nm and 20 ± 1 nm respectively, which is similar to the sizes measured with TEM.

Representative samples of the rod-like structures that contained encapsulated dsDNA were also analyzed using TEM. We started by the analysis of the rods formed without Mg2+ (Fig. S2). Elongated rods

with lengths exceeding 5μm and with a diameter of about 18 nm were observed. After optimization of the buffering conditions and upon the addition of Mg2+, we acquired well packed rods with reduced lengths,

up to 1μm (Fig. 3F-J). For all measured samples the diameters of the rods remain approximately 18 nm. We observed some polydispersity in the lengths of the rods, of which the average size seemed to depend on the length of dsDNA template used. Interestingly, we also observed some spherical structures with diameters ranging from 17 to 23 nm.

The rod-like VLPs with 100 bp dsDNA cargo have an average length of 90 nm and the particles appear oval. The VLPs with 200 bp dsDNA cargo, are already distinctly rod-like with an average length of 120 nm. The CP encasing 500 bp dsDNA formed rods with an average length of 180 nm, while the 1000 bp dsDNA mixed with CP assembles into rods with an average length of 300 nm and consequently high aspect ratios. There is an apparent correlation between the fully stretched length of the dsDNA scaffolds and the formed rod-like CP protein assemblies. dsDNA of 100 bp, 200 bp, 500 bp and 1000 bp has a theoretical length of 34 nm, 68 nm, 170 nm and 340 nm, respectively. Including the thickness of the CP end caps ( ± 5 nm), the predicted length of rod shaped VLPs is about 44, 78, 180 and 350 nm, which only slightly de-viates from the lengths measured by TEM (Fig. 3). The results of this section indicate that the assembly can indeed be controlled by the length of the dsDNA cargo, when the correct assembly conditions are used. This is beneficial in the creation of defined virus protein-based structures, for example in the fabrication of electronic nanowires.

The emergence of micron sized tube structures has been observed in literature. [32,41] Here, the formation of the tubes is attributed to the dsDNA being packaged staggered parallel to the tube axis, rather than coiled. Several strands of DNA could be packaged within the 10 nm diameter lumen of the tube to form salt bridges with the CP. The ob-served 18 nm diameter for CCMV-based rod-like structures is likely the result of the thermodynamically stable T = 1 type conformation of the CCMV capsid, which is in agreement with data reported by Zlotnick and his team [32]. The unexpected formation of spherical particles by dsDNA templating may be explained by interactions with the uranyl acetate staining agent [41] or the addition of Mg2+to the solution

(fewer spherical structures were observed in the Mg2+free samples,

Fig. S2), although to a lesser degree. In order to separate the structures based on geometry, we attempted several purification methods; i.e. SEC, sucrose gradient ultracentrifugation and cesium chloride ultra-centrifugation, however, these proved to be unsuccessful.

3.2. Cellular uptake route of native CCMV in different cell lines

Cellular uptake route of (nano)particles can be evaluated by using several well-studied inhibitors known to block certain pathways. The background on the different uptake mechanisms and the mode of action of the blocking agents are extensively discussed by Donaldson et al. [42]

Fig. 6. Uptake of different nanostructures of the protein capsid of CCMV in HeLa cells. Results and standard deviation are from mean fluorescent intensity average of 3 measurements fromflow cytometry after 4 h incubation. Only viable cells were used for analysis. (A) ATTO 647 N dye attached to (CP) the free proteins, (62 nt + CP - CCMV) spherical structures and (100 bp + CP - 1000 bp + CP) rod-like structures. (BeI) Show the uptake in percentage of different CCMV-based nanostructures with specific and nonspecific uptake inhibitors in HeLa cells. The nanostructures without inhibitor were normalized to 100% uptake. (B) Free CP dimers, (C) VLPs resulting from 62 nt ssDNA mixed with CP with Ø = 21 nm, (D) VLPs resulting from ssDNA of approximately 700 nt mixed with CP with Ø = 24 nm and (E) native CCMV with Ø = 28 nm, F) dsDNA of 100 bp mixed with CP, (G) dsDNA of 200 bp mixed with CP with average length of 120 nm, (H) dsDNA of 500 bp mixed with CP with average length of 180 nm and (I) dsDNA of 1000 bp mixed with CP with average length of 300 nm. *significant inhibition compared to CCMV (P < .05).

(9)

In our research Cytochalasin D is used to block phagocytosis and macropinocytosis; Fillipin III and Nystatin are used to specifically block caveolin dependent endocytosis; Dynasore to block both caveolin and clathrin dependent endocytosis; Chlorpromazine is applied to speci fi-cally block clathrin dependent endocytosis; and a hypertonic sucrose solution is applied to the cells to unspecifically block all uptake rou-tes.Fluorescently labeled polystyrene with a diameter of 1μm is used as

a positive control for phagocytosis, 10 kDa dextran for macro-pinocytosis, cholera toxin for caveolin dependent endocytosis and transferrin for clathrin dependent endocytosis(Fig. 4).

Different cell types have different functions, which results in dif-ferent dominant uptake pathways. Therefore, it is relevant to evaluate different cell lines to monitor the uptake route of CCMV into cells. The aim is to explore the possibilities of the medical use of CCMV-based

Fig. 7. Fluorescent confocal Images of paraformaldehydefixated Hela cells, after 4 h incubation with different CCMV-based nanostructures. The nucleus (DAPI) is stained in blue; early endosomes are stained in yellow (tagRFP-bacMam 2.0); and CPs are stained in red (ATTO 647 N). Showing: (A) free CPdimers, (B) 62 nt + CP, (C) 700 nt + CP, (D) Negative control, (E) 100 bp + CP, (F), 200 bp + CP, (G) 500 bp + CP and (H) 1000 bp + CP. (For interpretation of the references to colour in thisfigure legend, the reader is referred to the web version of this article.)

(10)

nanostructures. Cancer remains a prominent target for many therapies, therefore the HeLa cell line is chosen as thefirst cell type for evaluation. [43] For proper function in vivo, it is important for the particles to have long circulation times in the blood. A major hurdle in this case is the immune system, which is capable of degrading proteins and especially viral protein-based structures. Therefore, an immune cell line is eval-uated in the form of RAW 264.7 cells. As a third cell line Human Em-bryonic Kidney cells (HEK 293) are used, of which various uptake ex-periments are described. [44] Each cell type reacts differently to the inhibitors, so cell specific concentrations are optimized with the various positive controls for the selected uptake routes. The optimized con-centrations are shown in Table S1 and Fig. S3–5 which correspond to literature [40]. Fig. 5 shows the results of the ATTO 647 N labeled CCMV uptake in the different cell lines.

The uptake of the native CCMV in different cell lines is visualized in Fig. 5A. This shows no significant difference in uptake between the RAW 264.7 and HeLa cells, but a significantly lower (P < .01) uptake for HEK 293 cells. This is expected, because macrophages and certain cancer cells, with Hela as a prominent example, are known to have high uptake compared to HEK cells [43]. The effects of the different in-hibitors on the uptake of CCMV into these different cell lines is shown in Fig. 5B-D. In all of the used cell lines, filipin III, dynasore, chlor-promazine and cytochalasin D show significant inhibition (P < .01) of CCMV uptake, except for cytochalasin D in HEK293 cells.

This indicates that CCMV uses multiple uptake routes in the selected cell types and occurs through caveolae-mediated endocytosis, clathrin-mediated endocytosis and micropinocytosis. Fillipin III shows sig-nificant inhibition in all cell lines but Nystatin does not. This is un-expected because both compounds inhibit caveolin dependent en-docytosis as was confirmed by similar inhibition experiments with the positive control cholera toxin. These inhibitors alter the properties of cholesterol-rich membrane domains, but each in a different way [45]. The results apparently indicate that CCMV, uses caveolin mediated

endocytosis that is nystatin independent, which is also previously ob-served for the endocytosis of other viral structures [46].

Furthermore, in RAW264.7 cells (Fig. 5B) cytochalasin D shows significantly more (P < .05) inhibition compared to chlorpromazine, while this is reversed in HeLa cells (Fig. 5C). In HEK 293 (Fig. 5D) we measured no significant difference between either of the effective blocking agents. This would suggest that there is indeed a difference in uptake mechanisms in different cell lines. These differences mean that RAW264.7 cells mainly uses the macropinocytosis/phagocytosis pathway for CCMV uptake and that Hela cells primarily use clathrin-mediated endocytosis for CCMV uptake. The larger uptake inhibition in RAW264.7 cells with cytochalasin D is expected, because it is a mac-rophage, which is known to use phagocytosis for internalization of pathogens [47]. Interestingly, the results show that chlorpromazine is most effective in CCMV inhibition in Hela cells. This is different from the closely related CPMV, which used caveolin mediated endocytosis as the most prominent uptake route in HeLa cells [40]. Similarly, both viruses can utilize multiple uptake routes to enter the cells.

3.3. Shape dependent cellular uptake in HeLa cells

After revealing the uptake routes of the native CCMV in different cell lines, we studied the uptake pathways of the different nanos-tructures as described insection 2.12. For this we used HeLa cells be-cause this cell line is a relevant model system for cancer treatment. All nanostructures had the same degree of labelling with ATTO647 dye on their capsid proteins, allowing theirfluorescent intensity to be com-pared during uptake. The averagefluorescence intensity results of their uptake as monitored byflow cytometry are shown inFig. 6A.

Fig. 6A shows that the Ø = 21 nm (62 nt ssDNA) structure has the highest uptake (P < .05) and the Ø = 24 nm (700 nt ssDNA) structure the lowest uptake (P < .05) compared to the rest of the structures. As a next step, we determined the uptake route of the various nanostructures

Fig. 8. Fluorescent confocal images of paraformaldehydefixed Hela cells, with stained organelles, after 4 h incubation with CCMV. The nucleus (DAPI) is stained in blue; CCMV-CPs are stained in red (ATTO 647 N); (A) the early endosomes are stained in yellow bacMam 2.0); (B) late endosomes in yellow (tagRFP-bacMam 2.0); (C) lysosomes in yellow (tagRFP-(tagRFP-bacMam 2.0); (D) mitochondria in green (emGFP-(tagRFP-bacMam 2.0); (E) Golgi in green (emGFP-(tagRFP-bacMam 2.0); and (F) the cell membrane in blue (CF405S labeled Wheat Germ Agglutinin). (For interpretation of the references to colour in thisfigure legend, the reader is referred to the web version of this article.)

(11)

by evaluating the uptake in the presence of the different inhibitors (Fig. 6BeI).

The results show that the uptake of all the different nanostructures and free CP are significantly (P < .01) reduced by chlorpromazine. This inhibition has a similar degree of inhibition compared to the su-crose as a general inhibitor. Furthermore, for all nanostructures, except for 100 bp + CP (Fig. 6F), the reduction of the uptake by chlorproma-zine was significantly lower (P < .05) with respect to the filipin III, nystatin and cytochalasin D inhibitors. This indicates that clathrin-mediated endocytosis is the main uptake route that is utilized for all different nanostructures to enter HeLa cells. However, the uptake was not completely eliminated, which means that there are multiple path-ways possible.

An interesting difference in uptake, between the spherical and rod-like structures, is observed in the case offilipin III. For all the spherical structures including CP there is clear inhibition (P < .05), while most rod-like structures gave no significant inhibition (except for 1000 bp + CP). This suggests that the shape determines if caveolin mediated endocytosis of the particles occurs. Similar to previous results, the nystatin did not inhibit for all the nanostructures studied, con-firming that the uptake of CCMV-proteins occurs via nystatin in-dependent, caveolin mediated endocytosis. The results show that the larger rods (500 bp and 1000 bp) and the largest spherical particle (native CCMV) are inhibited (P < .05) better by cytochalasin D than the smaller spherical structures and rods (62 nt, 700 nt, 100 bp and 200 bp). This is expected, because macropinocytosis is in literature mostly observed for large particles, [48] however, the dimeric CPs are also inhibited by cytochalasin D. Upon analysis of these samples we observed that the free CPs partially aggregate in the medium and eventually precipitate out of the medium. This was not observed for the other CCMV-based nanostructures, which we confirmed by analyzing two representative nanostructures with TEM after 4 h incubation in the medium (Fig. S6). This aggregation could lead to larger structures, which can explain the inhibition by cytochalasin D.

The data discussed in the previous sections were obtained with CP labeled with ATTO 647 N dye. The exterior modification of particles can influence the uptake, which we considered by labelling on average only 1 out of 20 CPs. To further investigate the influence of the presence of the dye we also performed all experiments with an Alexafluor 647 dye coupled to the exterior lysines of the virus. The results show the same trend for the different inhibitors with clear (P < .01) uptake inhibition using chlorpromazine, which confirms that clathrin-medi-ated endocytosis is the main uptake route for all the used nanos-tructures. In order to completely exclude the influence of the dye modification of the CCMV CP-based particles, further studies with, for example, internal labelling are needed [23,49–51].

For all structures studied, it seems that clathrin-mediated en-docytosis is the primary route for uptake. Furthermore, it seems that there are multiple uptake routes possible in parallel, which show some shape dependence. This is different from previous studies investigating CPMV, which mainly followed caveolae-mediated endocytosis. Different geometries of TMV had different main uptake routes de-pending on the shape. [16,43] But, both work on TMV and our CCMV based VLPs show that the virus geometry has an influence on the cell entry pathway. The difference in uptake can be explained by the spe-cific protein chemistry on the exterior of the used viruses. Comparison between the TMV and CCMV studies indicate that the trends in shape dependent virus uptake are very dependent on the type of virus used and, so far, no general trend applicable for all virus structures is ob-served.

3.4. Intracellular positioning of CCMV-based nanostructures

The results from the previous section indicate that clathrin-en-docytosis is the general uptake route for the nanostructures in HeLa cells. To verify this observation, we analyzed the cells usingfluorescent

confocal microscopy by staining the organelles with BacMam 2.0 CellLight reagents. We started with the early endosomes to monitor the uptake. The BacMam-early endosome uses a gene transfection encoding for a Rab5a protein fused to a tagRFP, which is delivered to the cells using the baculovirus insect virus. Rab5a is a natural enzyme that lo-calizes in the early endosomes [52]. The results depicted inFig. 7show that there is colocalization of the capsid proteins of the CCMV-based nanostructures with the early endosomes in the cells, as indicated by overlapping orange spots. All structures, except for the CP, show a moderate Pearson correlation with an average of 0.6 ± 0.2 for the RFP with the ATTO 647 N dyes on the capsid proteins confirming that the uptake of the particles is indeed mediated by endocytosis.

We observed an average negative Pearson correlation (−0.3 ± 0.2) with the DAPI blue staining in healthy looking cells for all the nanostructures. This indicates that the CPs do not enter the nucleus, but this does not mean that the DNA can't. Interestingly, all the nanostructures show a wide distribution offluorescent spots in the cell, except for the free CP dimers. These show more clusteredfluorescence, which is indicative for aggregation in the cell medium, leading to a low correlation of 0.1 ± 0.3 with the endosomes. Literature reports state that free CPs are less stable in their dimeric shape compared to the assembled one, which corresponds to our observations in cells. This aggregation behavior was also discussed in the previous section and further explains the higher inhibition by cytochalasin D.

We continued with the native CCMV to determine in which orga-nelles the plant virus particles end up (Fig. 8). We again used the BacMam 2.0 CellLight system to colour the organelles. The late endo-somes are visualized using Rab7a linked to tagRFP; the lysoendo-somes by lamp1 (lysosomal associated membrane protein 1) fused to a tagRFP; the mitochondria by emGFP linked to a mitochondria residing enzyme (E1 alpha pyruvate dehydrogenase); and the Golgi apparatus by emGFP linked to a Golgi resident enzyme (N-acetylgalactosaminyltransferase). The results are quantified using the Pearson coefficient, which show a weak positive correlation for the colocalization of the CCMV proteins in the mitochondria (0.2 ± 0.1), a low correlation for the Golgi (0.35 ± 0.2), a medium to strong correlation with the lysosomes and membrane (both 0.6 ± 0.2) and a strong correlation in the early and late endosomes (0.7 ± 0.2 and 0.75 ± 0.1, respectively). The results compare well to studies with the closely related CPMV [44]. For this virus some minor colocalization in the Golgi was reported, which we also observed. Furthermore, a clear colocalization of the CCMV in the endosomes is shown, confirming the endocytosis pathway.Fig. 8F also shows red CPs attached to the cell exterior, on the membrane, in-dicating that not all virus particles are taken up. Colocalization is also found in the lysosomes, since this correlation is not as high as with the endosomes it appears that some of the viruses escape the lysosomes. 4. Conclusions

In this contribution different lengths of ssDNA and dsDNA were encapsulated using CCMV-CPs, creating virus-like structures of various shapes and sizes at physiological conditions. We evaluated the in vitro uptake route and intracellular fate of these CCMV-based nanos-tructures.

Encapsulation of ssDNA resulted in VLPs with a spherical shape, where increasing the length of the ssDNA caused an apparent increase in the particle diameter. For the dsDNA encapsulation, the addition of the magnesium ions to the assembly buffer reduced the length of the formed rod-like structures and resulted in lengths corresponding to the stretched dsDNA template length.

We have shown that macrophages (RAW264.7) and cancer (HeLa) cells have a higher uptake efficiency for the native CCMV than in epi-thelial (HEK 293) cells. For all 3 cell lines and CCMV-based nanos-tructures, it seems that multiple uptake pathways are possible. There is an indication that the dominant uptake pathway is dependent on the cell line. In HeLa cells, all of the different nanostructures use

(12)

mediated endocytosis as the main uptake route. The data also give some indication that the larger structures show an increased utilization of macropinocytosis, while the spherical CCMV-based VLPs have the tendency to use caveolin mediated endocytosis.

After 4 h of incubation, most of the nanostructures are localized in the endosomes, lysosomes or on the membranes. Scarcely localization is observed in the nucleus, Golgi and mitochondria. Different virus-based structures showed minimal differences in their cell localization, except for the free CPs, which aggregated in the medium.

Overall, the presented work gives us a better understanding of how CCMV and CCMV-based nanostructures interact with cells, pointing to a shape dependence in the cellular uptake process. This brings us one step closer to understanding why so many different viral shapes exist in nature and how this can lead to the design of better virus-based na-nomedicine, in particular for the potential application of CCMV-CP based VLPs in therapeutics and diagnostics.

Acknowledgements

The authors thank C. Breukers for his help with theflow cytometry, I.B.M. Siemerink-Konings for her help with the confocal microscope and P.H. Hamming for the Matlab scripts for the processing of the cell data. This work isfinancially supported by the European Research Council (Consolidator Grant‘Protcage’ #616907).

Appendix A. Supplementary data

Supplementary data to this article can be found online athttps:// doi.org/10.1016/j.jconrel.2019.06.019.

References

[1] B.D. Lindenbach, M.J. Evans, A.J. Syder, B. Wölk, T.L. Tellinghuisen, C.C. Liu, T. Maruyama, R.O. Hynes, D.R. Burton, J.A. McKeating, C.M. Rice, Complete re-plication of hepatitis C virus in cell culture, Science 309 (2005) 623–626. [2] S.J. Maassen, M.V. de Ruiter, S. Lindhoud, J.J.L.M. Cornelissen, Oligonucleotide

length-dependent formation of virus-like particles, Chem. Eur. J. 24 (2018) 7456–7463.

[3] P. Lam, N.F. Steinmetz, Plant viral and bacteriophage delivery of nucleic acid therapeutics, Wiley Interdisciplinary Reviews: Nanomedicine and

Nanobiotechnology 10 (2018) e1487.

[4] K.J. Koudelka, A.S. Pitek, M. Manchester, N.F. Steinmetz, Virus-based nanoparticles as versatile Nanomachines, Annual review of Virology 2 (2015) 379–401. [5] L. Yang, A. Liu, M.V. de Ruiter, C.A. Hommersom, N. Katsonis, P. Jonkheijm,

J.J.L.M. Cornelissen, Compartmentalized supramolecular hydrogels based on viral nanocages towards sophisticated cargo administration, Nanoscale 10 (2018) 4123–4129.

[6] O. Azizgolshani, R.F. Garmann, R. Cadena-Nava, C.M. Knobler, W.M. Gelbart, Reconstituted plant viral capsids can release genes to mammalian cells, Virology 441 (2013) 12–17.

[7] D. Blaas, Viral entry pathways: the example of common cold viruses, Wien. Med. Wochenschr. 166 (2016) (1946) 211–226.

[8] A. Akinc, G. Battaglia, Exploiting endocytosis for nanomedicines, Cold Spring Harb. Perspect. Biol. 5 (2013) a016980.

[9] C.M. Beddoes, C.P. Case, W.H. Briscoe, Understanding nanoparticle cellular entry: a physicochemical perspective, Adv. Colloid Interf. Sci. 218 (2015) 48–68. [10] T.-G. Iversen, T. Skotland, K. Sandvig, Endocytosis and intracellular transport of

nanoparticles: present knowledge and need for future studies, Nano Today 6 (2011) 176–185.

[11] X.-Y. Sun, Q.-Z. Gan, J.-M. Ouyang, Size-dependent cellular uptake mechanism and cytotoxicity toward calcium oxalate on Vero cells, Sci. Rep. 7 (2017) 41949. [12] E. Fröhlich, The role of surface charge in cellular uptake and cytotoxicity of medical

nanoparticles, Int. J. Nanomedicine 7 (2012) 5577–5591.

[13] B.D. Chithrani, A.A. Ghazani, W.C.W. Chan, Determining the size and shape de-pendence of gold nanoparticle uptake into mammalian cells, Nano Lett. 6 (2006) 662–668.

[14] J. Sun, L. Zhang, J. Wang, Q. Feng, D. Liu, Q. Yin, D. Xu, Y. Wei, B. Ding, X. Shi, X. Jiang, Tunable rigidity of (polymeric Core)–(lipid Shell) nanoparticles for regulated cellular uptake, Adv. Mater. 27 (2015) 1402–1407.

[15] F. Zhao, Y. Zhao, Y. Liu, X. Chang, C. Chen, Y. Zhao, Cellular uptake, intracellular trafficking, and cytotoxicity of nanomaterials, Small 7 (2011) 1322–1337. [16] Q. Wei, C. Huang, Y. Zhang, T. Zhao, P. Zhao, P. Butler, S. Zhang,

Mechanotargeting: mechanics-dependent cellular uptake of nanoparticles, Adv. Mater. 30 (2018) 1707464.

[17] X. Liu, F. Wu, Y. Tian, M. Wu, Q. Zhou, S. Jiang, Z. Niu, Size dependent cellular uptake of rod-like bionanoparticles with different aspect ratios, Sci. Rep. 6 (2016)

24567.

[18] S. Shukla, F.J. Eber, A.S. Nagarajan, N.A. Difranco, N. Schmidt, A.M. Wen, S. Eiben, R.M. Twyman, C. Wege, N.F. Steinmetz, The impact of aspect ratio on the biodis-tribution and tumor homing of rigid soft-matter nanorods, Advanced Healthcare Materials 4 (2015) 874–882.

[19] J.B. Bancroft, E. Hiebert, Formation of an infectious nucleoprotein from protein and nucleic acid isolated from a small spherical virus, Virology 32 (1967) 354–356. [20] L. Lavelle, J.P. Michel, M. Gingery, The disassembly, reassembly and stability of

CCMV protein capsids, J. Virol. Methods 146 (2007) 311–316.

[21] L.O. Liepold, J. Revis, M. Allen, L. Oltrogge, M. Young, T. Douglas, Structural transitions in cowpea chlorotic mottle virus (CCMV), Phys. Biol. 2 (2005) S166–S172.

[22] A. Liu, M. Verwegen, M.V. De Ruiter, S.J. Maassen, C.H.H. Traulsen,

J.J.L.M. Cornelissen, Protein cages as Containers for Gold Nanoparticles, J. Phys. Chem. B 120 (2016) 6352–6357.

[23] O. Tagit, M.V. De Ruiter, M. Brasch, Y. Ma, J.J.L.M. Cornelissen, Quantum dot encapsulation in virus-like particles with tuneable structural properties and low toxicity, RSC Adv. 7 (2017) 38110–38118.

[24] S.J. Maassen, A.M. Van Der Ham, J.J.L.M. Cornelissen, Combining protein cages and polymers: from understanding self-assembly to functional materials, ACS Macro Lett. 5 (2016) 987–994.

[25] A. Arnida, H. Ghandeharia Malugin, Cellular uptake and toxicity of gold-nanoparticles in prostate cancer cells: a comparative study of rods and spheres, Journal of Applied Toxicology 30 (2010) 212–217.

[26] M. Kwak, I.J. Minten, D.-M. Anaya, A.J. Musser, M. Brasch, R.J.M. Nolte, K. Müllen, J.J.L.M. Cornelissen, A. Herrmann, Virus-like Particles Templated by DNA Micelles: A General Method for Loading Virus Nanocarriers, J. Am. Chem. Soc. 132 (2010) 7834–7835.

[27] J.G. Millán, M. Brasch, E. Anaya-Plaza, A. de la Escosura, A.H. Velders, D.N. Reinhoudt, T. Torres, M.S.T. Koay, J.J.L.M. Cornelissen, Self-assembly trig-gered by self-assembly: Optically active, paramagnetic micelles encapsulated in protein cage nanoparticles, Journal of Inorganic Biochemistry 136 (2014) 140–146. [28] S. Sinn, L. Yang, F. Biedermann, D. Wang, C. Kübel, J.J.L.M. Cornelissen, L. De Cola,

Templated formation of luminescent virus-like particles by tailor-made Pt(II) Amphiphiles, J. Am. Chem. Soc. 140 (2018) 2355–2362.

[29] Y. Wu, H. Yang, H.-J. Shin, Encapsulation and crystallization of Prussian blue na-noparticles by cowpea chlorotic mottle virus capsids, Biotechnol. Lett. 36 (2014) 515–521.

[30] D. Luque, A. de la Escosura, J. Snijder, M. Brasch, R.J. Rose, M.S.T. Koay, J.L. Carrascosa, G.J.L. Wuite, W.H. Roos, A.J.R. Heck, J.J.L.M. Cornelissen, T. Torres, J.R. Caston, Self-assembly and characterization of small and mono-disperse dye nanospheres in a protein cag, Chem. Sci. 5 (2014) 575–581. [31] J.B. Bancroft, K.M. Smith, M.A. Lauffer, F.B. Bang (Eds.), The Self-Assembly of

Spherical Plant Viruses, Academic Press, 1970, pp. 99–134. Advances in Virus Research.

[32] S. Mukherjee, C.M. Pfeifer, J.M. Johnson, J. Liu, A. Zlotnick, Redirecting the coat protein of a spherical virus to assemble into tubular nanostructures, J. Am. Chem. Soc. 128 (2006) 2538–2539.

[33] L. Schoonen, J. Pille, A. Borrmann, R.J.M. Nolte, J.C.M. Van Hest, Sortase A-mediated N-terminal modification of cowpea chlorotic mottle virus for highly ef-ficient cargo loading, Bioconjug. Chem. 26 (2015) 2429–2434.

[34] M. Brasch, R.M. Putri, M.V. De Ruiter, D. Luque, M.S.T. Koay, J.R. Castón, J.J.L.M. Cornelissen, Assembling enzymatic cascade pathways inside virus-based nanocages using dual-tasking nucleic acid tags, J. Am. Chem. Soc. 139 (2017) 1512–1519.

[35] M. Comellas-Aragones, H. Engelkamp, V.I. Claessen, N.A.J.M. Sommerdijk, A.E. Rowan, P.C.M. Christianen, J.C. Maan, B.J.M. Verduin, J.J.L.M. Cornelissen, R.J.M. Nolte, A virus-based single-enzyme nanoreactor, Nat. Nanotechnol. 2 (2007) 635–639.

[36] L. Sánchez-Sánchez, R.D. Cadena-Nava, L.A. Palomares, J. Ruiz-Garcia, M.S.T. Koay, J.J.M.T. Cornelissen, R. Vazquez-Duhalt, Chemotherapy pro-drug ac-tivation by biocatalytic virus-like nanoparticles containing cytochrome P450, Enzym. Microb. Technol. 60 (2014) 24–31.

[37] R.F. Garmann, M. ComGarcia, A. Gopal, C.M. Knobler, W.M. Gelbart, The as-sembly pathway of an icosahedral single-stranded RNA virus depends on the strength of inter-subunit attractions, J. Mol. Biol. 426 (2014) 1050–1060. [38] J. Mikkilä, A.P. Eskelinen, E.H. Niemelä, V. Linko, M.J. Frilander, P. Törmä,

M.A. Kostiainen, Virus-encapsulated DNA origami nanostructures for cellular de-livery, Nano Lett. 14 (2014) 2196–2200.

[39] R. Konecny, J. Trylska, F. Tama, D. Zhang, N.A. Baker, C.L. Brooks,

J.A. McCammon, Electrostatic properties of cowpea chlorotic mottle virus and cu-cumber mosaic virus capsids, Biopolymers 82 (2006) 106–120.

[40] E.M. Plummer, M. Manchester, Endocytic uptake pathways utilized by CPMV na-noparticles, Mol. Pharm. 10 (2013) 26–32.

[41] K. Burns, S. Mukherjee, T. Keef, J.M. Johnson, A. Zlotnick, Altering the energy landscape of virus self-assembly to generate kinetically trapped nanoparticles, Biomacromolecules 11 (2010) 439–442.

[42] D. Dutta, J.G. Donaldson, Search for inhibitors of endocytosis: intended specificity and unintended consequences, Cellular Logistics 2 (2012) 203–208.

[43] J.R. Masters, HeLa cells 50 years on: the good, the bad and the ugly, Nat. Rev. Cancer 2 (2002) 315.

[44] P. Thomas, T.G. Smart, HEK293 cell line: a vehicle for the expression of re-combinant proteins, J. Pharmacol. Toxicol. Methods 51 (2005) 187–200. [45] P. Lajoie, I.R. Nabi, Chapter 3 - lipid rafts, Caveolae, and their endocytosis, in:

K.W. Jeon (Ed.), International Review of Cell and Molecular Biology, Academic Press, 2010, pp. 135–163.

(13)

[46] N.-C. I., E. A., N. S., P.d.L. M.C, Caveolae as an additional route for influenza virus endocytosis in MDCK cell, Cell. Mol. Biol. Lett. 9 (2004) 47–60.

[47] David M. Richards, Robert G. Endres, The mechanism of phagocytosis: two stages of engulfment, Biophys. J. 107 (2014) 1542–1553.

[48] A. Muthuraman, Chapter 3 - the current perspectives of nanoparticles in cellular and organ-specific drug targeting in biological system, in: A.M. Grumezescu (Ed.), Nanostructures for the Engineering of Cells, Tissues and Organs, William Andrew Publishing, 2018, pp. 105–154.

[49] J.M. Hooker, E.W. Kovacs, M.B. Francis, Interior surface modification of

bacteriophage MS2, J. Am. Chem. Soc. 126 (2004) 3718–3719.

[50] F. Li, Z.-P. Zhang, J. Peng, Z.-Q. Cui, D.-W. Pang, K. Li, H.-P. Wei, Y.-F. Zhou, J.-K. Wen, X.-E. Zhang, Imaging viral behavior in mammalian cells with self-as-sembled capsid–quantum-dot hybrid particles, Small 5 (2009) 718–726. [51] M.V. de Ruiter, N.J. Overeem, G. Singhai, J.J.L.M. Cornelissen, Induced Förster

resonance energy transfer by encapsulation of DNA-scaffold based probes inside a plant virus based protein cage, J. Phys. Condens. Matter 30 (2018) 184002. [52] J. Huotari, A. Helenius, Endosome maturation, EMBO J. 30 (2011) 3481–3500.

Referenties

GERELATEERDE DOCUMENTEN

De ontwikkelingen in de ons om- ringende landen zijn bepalend voor de Nederlandse graanteelt, maar de daar ontwikkelde kennis kan echter niet zomaar worden in- gekocht of

De politie uit Mont-de-Marsan, onder leiding van kapitein Jourdren, begon ’s avonds om 11 uur te zoeken, waar-.. schuwde de leiding van de groeve, die ’s nachts afgesloten is, en

Verloop van de krachten in drie ankerbouten in de verankering van element 36 tegen de tijd tijdens de aanrijding met de 30 tons vrachtauto tegen de

Maar al te vaak zijn deze boeken geschreven voor grate onder- nemingen en niet direkt toepasbaar voor het MKB. In de twee nog te verschijnen artikelen zullen

organisatiecultuur bestaan. 224) cultuur als “de collectieve mentale programmering die de leden van de organisatie onderscheidt van die van een andere”. Hieruit blijkt dat cultuur

Results of table 4.10 show a significant simple main effect of health consciousness in the unhealthy prime condition on sugar and saturated fat content of baskets,

It consists of four different positive sense RNAs encapsulated in three capsids of the same structure, made from 90 homodimers of the capsid proteins CP, which form the protein