Cytocompatibility and Cellular Internalization of ... - ACS Publications

Apr 26, 2018 - ABSTRACT: The recently developed synthetic oligonucleo- tides referred to as “click” nucleic acids (CNAs) are promising due to thei...
0 downloads 0 Views 2MB Size
Article Cite This: Biomacromolecules XXXX, XXX, XXX−XXX

pubs.acs.org/Biomac

Cytocompatibility and Cellular Internalization of PEGylated “Clickable” Nucleic Acid Oligomers Alex J. Anderson,† Erica B. Peters,† Alex Neumann,† Justine Wagner,† Benjamin Fairbanks,† Stephanie J. Bryant,*,†,‡,§ and Christopher N. Bowman†,‡,§ †

Department of Chemical and Biological Engineering, ‡Material Science and Engineering Program, and §BioFrontiers Institute, University of Colorado, Boulder, Colorado 80303, United States S Supporting Information *

ABSTRACT: The recently developed synthetic oligonucleotides referred to as “click” nucleic acids (CNAs) are promising due to their relatively simple synthesis based on thiol−X reactions with numerous potential applications in biotechnology, biodetection, gene silencing, and drug delivery. Here, the cytocompatibility and cellular uptake of rhodamine tagged, PEGylated CNA copolymers (PEG-CNA-RHO) were evaluated. NIH 3T3 fibroblast cells treated for 1 h with 1, 10, or 100 μg/mL PEG-CNA-RHO maintained an average cell viability of 86%, which was not significantly different from the untreated control. Cellular uptake of PEG-CNA-RHO was detected within 30 s, and the amount internalized increased over the course of 1 h. Moreover, these copolymers were internalized within cells to a higher degree than controls consisting of either rhodamine tagged PEG or the rhodamine alone. Uptake was not affected by temperature (i.e., 4 or 37 °C), suggesting a passive uptake mechanism. Subcellular colocalization analysis failed to indicate significant correlations between the internalized PEGCNA-RHO and the organelles examined (mitochondria, endoplasmic reticulum, endosomes and lysosomes). These results indicate that CNA copolymers are cytocompatible and are readily internalized by cells, supporting the idea that CNAs are a promising alternative to DNA in antisense therapy applications.



of action, often requiring the use of cytotoxic polymers,7 and once the oligonucleotides are released into the cytosol, they are vulnerable to intracellular degradation.8,9 To prevent oligonucleotide degradation, the nucleotide structure can be chemically modified such as in phosphorothioate nucleic acids (PS-DNA) or peptide nucleic acids (PNAs). By altering or eliminating the phosphodiester bond, PS-DNA, PNAs, and other types of modified nucleic acids have proven to be nuclease resistant.10−12 In some cases, they have been shown to have higher transfection efficiencies.13 However, like DNA synthesis, the assembly of most modified nucleic acids oligomers relies on a complicated solid-phase strategy,14,15 which is low yield and difficult to purify, time-consuming, and can require the use of hazardous chemicals and solvents.16 In contrast, recently reported “click” nucleic acids (CNAs) are nucleic acid oligomers synthesized by taking advantage of thiol-mediated click reactions.17 CNA monomers bare a resemblance to DNA nucleotides in that they contain a pendent nucleobase; however, they differ in that the backbone contains a thioether structure that arises from the thiol−X reactions that are used to form the CNA. By employing conditions that favor the correct thiol−X reaction, carefully

INTRODUCTION Nucleic acids represent one of the most powerful, functional biological materials due to their ability to control and direct biological function at all levels and to transmit genetic information from one generation to the next. These functions are possible solely because of the specific nucleotide sequence of each polymer and the nucleobase’s ability to recognize and hybridize with its complementary base. These characteristics are what has made nucleic acids a popular topic in both material science and biotechnology. Specifically, the ability of DNA to recognize and bind to highly specific genetic sequences has led to its use as antisense agents in gene therapy. Significant reductions in protein expression, and thereby a modulation of disease symptoms, have been achieved by delivering oligomeric DNA that is complementary to targeted sequences of either DNA or mRNA.1−3 While gene therapy has been a major area of research,4,5 the field has largely been hindered by the inefficient cellular delivery of DNA, instability of foreign DNA in biological environments, and the limited scale of DNA synthesis. Several techniques have been developed to circumvent these shortcomings. One common strategy is to condense antisense oligonucleotides into complexed nanoparticle delivery systems.6 This strategy has been shown to increase transfection efficiency as well as protect them from extracellular degradation. However, these nanoparticles have a complicated mechanism © XXXX American Chemical Society

Received: January 31, 2018 Revised: April 11, 2018 Published: April 26, 2018 A

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

Cell Culture. NIH 3T3 fibroblast cells obtained from ATCC were cultured in DMEM growth media supplemented with 10% FBS and 1% antibiotics at 37 °C and 5% CO2. Cells were passaged every 3−5 days or when 90% confluency was reached with 0.5% trypsin−EDTA and replating at a 1:10 ratio. Cell Viability. Viability was assessed by an MTT assay. Briefly, cells were seeded in a 96 well plate at 800 cells/well and allowed to attach and grow for 48 h. Cells were then treated with varying concentrations of PEG-CNA-RHO and PEG-RHO conjugate (0−100 μg/mL) in triplicate for 1 h. A positive control was performed by adding a small punch of a powdered latex glove to the wells. After treatment, cells were washed with PBS and cultured in fresh media for another 48 h. The assay was performed by replacing the media in each well with 100 μL of fresh media and adding 10 μL of 12 mM MTT to each well after which the plate was incubated at 37 °C and 5% CO2 for 4−6 h. The resulting crystal precipitates were dissolved by adding 100 μL of 0.1 g/ mL sodium dodecyl sulfate in 0.1 M HCl, after which the solution’s absorbance was measured at 570 nm. Cellular Uptake of Fluorescent Conjugates. Cells were plated in a high quality, glass bottom 96-well plate (ibidi μ-Plate) at a concentration of 4000−7000 cells/well and allowed to adhere and grow overnight at 37 °C and 5% CO2. The culture medium was replaced with 135 μL of fresh media, to which 15 μL of the 1 mg/mL PEG-CNA-RHO solution was added. The plate was swirled gently to allow mixing. After incubation for 1 h, the cells were washed with PBS, and fresh media was added to the wells. Time-Dependent and Temperature-Dependent Cellular Uptake. For time-dependent studies, cells were plated and treated with PEG-CNA-RHO as described previously, but with varying incubation times from 30 s to 1 h. After incubation, the cells were washed with PBS and immediately fixed with 4% paraformaldehyde for 10 min at room temperature. After washing with PBS, cells were incubated with fresh PBS overnight at 4 °C, protected from light. For temperature-dependent studies, cells were plated as described previously but were incubated at either 4 or 37 °C prior to treatment with PEG-CNA-RHO. Cells were incubated with PEG-CNA-RHO at 4 or 37 °C for 1 h, after which the cells were washed with PBS and fixed with 4% paraformaldehyde. Immunofluorescence. Cells were plated and treated with PEGCNA-RHO the same way as before. After treatment, fresh media was added to the wells, and the cells were immediately imaged to verify uptake. Cells were then fixed with 4% paraformaldehyde for 20 min at room temperature, after which they were washed twice with PBS. Cells were permeabilized with 0.5% Triton X-100 for 5 min at room temperature, washed twice with PBS, and then blocked with 3% BSA for 60 min at room temperature. After rinsing with PBS, cells were stained with primary antibodies (endoplasmic reticulum:calreticulin antibody, rabbit anti-mouse, Novus Biologicals, used at 1:90 dilution in 3% BSA; mitochondria: ATPB antibody, mouse monoclonal, Abcam, used at 1:500 dilution in 3% BSA; endosome: EEA1 antibody, mouse monoclonal, R&D Systems , used at 1:50 dilution in 3% BSA; lysosome: LAMP2B antibody, rabbit anti-mouse, Abcam, used at 1:600 dilution in 3% BSA) for 1 h at room temperature. After rinsing wells three times with PBS, all traces of the primary antibody were removed by incubating with 0.01% Tween-20 for 1 h at room temperature. The cells were washed with PBS twice and then stained with the fluorescent secondary antibody (AlexaFluor 488, donkey antimouse, Invitrogen, used at 1:200 dilution in 3% BSA; AlexaFluor 647 donkey anti-rabbit, Invitrogen, used at 1:200 dilution in 3% BSA) for 1 h at room temperature in the dark. After washing with PBS three times, F-actin was stained with and phallotoxins (Phalloidin 488 and Phalloidin 647, Invitrogen, used at 1:60 dilution in 3% BSA) for 20 min at room temperature. Cells were washed with PBS twice and then stained with Hoechst 33342 (Invitrogen, used at 1 μg/mL in 3% BSA) for 5 min at room temperature. Finally, cells were washed with PBS twice and stored in PBS at 4 °C until imaged. Fluorescence Microscopy. Cells with fluorescent components were imaged on a Nikon Spinning Disc confocal microscope. To evaluate the extent of cellular uptake, a 40× air objective was used. For single cell images, a 100× oil immersion objective was used. For all

designed CNA monomers are polymerized into CNA polymer chains quickly and under ambient settings, making them an interesting and attractive alternative form of oligomeric nucleobases. CNA structure eliminates the phosphodiester bond, rendering it invisible to both endo- and exonucleases.18,19 Furthermore, while complete specific and arbitrary sequence control is still out of reach, sets of orthogonal thiol−X reactions can be used to build nucleic acid sequences that are polymerized in a single step to yield repetitive sequences of nucleotides. Thus, CNAs have the potential to be used for antisense gene therapy strategies, either as the antisense agent itself for treatment of diseases caused by repeat expansions20 or as a complexation agent for DNA delivery.21 In addition, they may offer many advantages over current antisense agents, such as nuclease resistance and a more facile synthesis. However, before CNAs can be realized for such applications, it is necessary to understand whether they will be compatible with biological systems. This study investigated cytocompatibility, cell uptake properties, and interactions with subcellular organelles of CNAs in vitro. Copolymers of CNA and poly(ethylene glycol) (PEG) were synthesized to create water-soluble CNA conjugates. PEGylated CNAs functionalized with a fluorophore were incubated with cells, and cytocompatibility was determined by measuring cell metabolism. Cellular internalization was assessed by visualizing uptake with fluorescence microscopy. Finally, immunofluorescent techniques were employed to study subcellular colocalization of CNA copolymers upon internalization.



EXPERIMENTAL SECTION

PEG-CNA-RHO and PEG-RHO Conjugate Synthesis. The thymine CNA monomer was synthesized as previously described.16 Photoinitiated copolymerization was carried out by dissolving thiolated poly(ethylene glycol) (PEG-SH) (MW 2000 g/mol, Sigma-Aldrich) at 100 mM and thymine CNA monomer at 1 M in DMF containing 0.01 wt % 2,2-dimethoxy-2-phenylacetophenone (DMPA). After irradiation with 365 nm light for 15 min, the resulting material was precipitated into cold diethyl ether. The supernatant was decanted leaving the solid copolymer (PEG-CNA), which was dried and resuspended in DI H2O. After agitating the polymer suspension overnight, the insoluble polymer fraction was spun down, and the supernatant was collected and lyophilized, yielding water-soluble PEGCNA whose presence was confirmed by 1H NMR (Bruker AV-III 400 MHz) and GPC (Tosoh HLC-8320GPC). In all cases, GPC samples were run in DMSO using two detectors: refractive index (RI) and ultraviolet (UV). Average molecular weights were calculated using PMMA standards. PEG-CNA was then dissolved at 500 μM in PBS in a scintillation vial, and disulfides were reduced with tris(2carboxyethyl)phosphine (TCEP) (Chem-Impex International) for 20 min while the solution was purged with inert gas. Then, under an inert atmosphere, 2.5 equiv of Rhodamine Red (RHO) C2 maleimide dissolved in DMSO (7.4 mM) was added to the stirring PEG-CNA solution and allowed to react for 2−3 h in the dark. The reaction solution was then dialyzed in the dark with a 1 kDa MWCO against DI H2O for 2 days (until the diasylate appeared colorless), yielding a pink fluorescent solution. The dialyzed PEG-CNA-RHO was then lyophilized and analyzed by 1H NMR and GPC. The extent of labeling by the dye was determined by UV−vis. The lyophilized PEGCNA-RHO powder was then reconstituted at a concentration of 1 mg/mL in PBS. This solution was sterile filtered with a 0.22 μm, PVDF, sterile syringe filter and frozen in aliquots. PEG-RHO was synthesized in a similar manner by reacting PEG-SH (MW 2000 g/ mol) with Rhodamine Red C2 maleimide. Once the fluorescent molecule was introduced, care was taken to protect the compound from exposure to light. B

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules cases, multiple images across multiple replicate wells were taken. Reported images are representative of the corresponding experimental conditions. All image analysis was completed with the ImageJ software. Statistical Analysis. Data presented are means of replicates (n, indicated in figure captions) with error bars representing standard deviations. Statistical analysis for the MTT assay and cell uptake assay was performed using a One-Way ANOVA with respect to cell viability and average fluorescence intensity per cell, respectively. Statistical significance was defined at the 95% confidence limit (p < 0.05). In cases where significant differences were present, comparisons were made using Tukey’s post-hoc analysis to identify differences between groups.

Scheme 1. In Vitro Strategy for Evaluating the Cytocompatibility of PEGylated CNA Copolymers



RESULTS AND DISCUSSION To begin evaluating CNAs’ cytocompatibility, linear, PEGylated CNAs were synthesized to increase the hydrophilicity of the CNA chain. Because of the lack of charge and hydrophilic groups on the CNA backbone, CNA homopolymers exhibit poor water solubility. It has also been shown that the addition of PEG to biomaterials can help increase the material’s bioavailability,22 protect the material from degradation,23 and stabilize nucleic acid hybridization interactions.24 Photoinitiated copolymerization of a thiolated PEG (molecular weight of 2000 g/mol) and thymine CNA monomers yielded a linear block copolymer named PEG-CNA. This type of copolymerization was chosen as the preferred method of synthesis because it is a simple, one-step reaction and because it results in a thiol end group on the CNA block which can subsequently be functionalized with a fluorophore. Gel permeation chromatography of this copolymer showed an increase in molecular weight from the PEG-SH, suggesting successful copolymerization, along with a low-molecular-weight peak that is attributed to the cyclized monomer (Supporting Information Figure S1). The water-soluble fraction of the block copolymer was isolated, and GPC and 1H NMR analysis confirmed that this copolymer contained an average of five thymine repeat units per chain (as obtained by peak integration with respect to the PEG methoxy peak) (see Figure S5). It should be noted that the degree of polymerization calculated here is less than our original molar excess of thymine monomer (10:1, thymine monomer:PEGSH). This behavior is because the cyclized monomer that is observed in the GPC trace, unreacted CNA homopolymers, and copolymers with higher degrees of polymerization are less likely to be solubilized by the water due to their high CNA content. Finally, a fluorophore (Rhodamine Red C2 maleimide) was conjugated to the sulfhydryl-terminated end of the CNA chain to yield the PEG-CNA-RHO conjugate. Effective conjugation was confirmed by 1H NMR analysis and UV−vis (Figures S6 and S7). The GPC trace of the resulting compound showed that the major peak is maintained and that many lowmolecular-weight peaks from the initial copolymerization disappear (Figure S8). Experiments were conducted using the strategy summarized in Scheme 1. Internalization experiments were performed by incubating cells with known concentrations of PEG-CNARHO, fluorescent PEG (PEG-RHO), or rhodamine dye alone (RHO), followed by subsequent washing. In this study, NIH 3T3 fibroblast cells were employed, which have been used previously to investigate cytocompatibility and cell uptake.25−27 The cytocompatibility of PEG-CNA-RHO was determined by the MTT assay, which measures cellular metabolism. Since metabolic activity is directly proportional to cell number, the MTT assay is an indirect measurement of cell viability or how

many cells survive after being treated with a molecule of interest. Concentrations of PEG-CNA-RHO ranging from 0 to 100 μg/mL were incubated with NIH 3T3 cells for a period of 1 h. Over the concentration range tested, cell metabolic activity was not statistically different from the untreated control (Figure 1), indicating cytocompatibility. As a positive control, cells were

Figure 1. An MTT assay confirms the low toxicity of the synthetic conjugates delivered in this study. Tukey’s post hoc comparison failed to identify significant differences in viability at a 95% confidence limit. (n = 3, error bars represent standard deviations.)

incubated with a powdered latex glove for the same amount of time, which reduced cell metabolism to ∼20%.28 Cells were also treated with PEG-RHO in the same concentration range as above. Cell viability after treatment with PEG-CNA-RHO was not statistically different from the viability of cells treated with PEG-RHO, which indicates that the CNA block does not add any significant toxicity. To examine the effect of the PEG-CNA-RHO copolymer structure on cell uptake, cells were incubated with PEG-CNARHO, fluorescent PEG (PEG-RHO), and the fluorescent dye alone (RHO). Figures 2 shows the extent of cellular uptake of representative samples after treatment with the three types of molecules. The bright red fluorescent spots seen in the first panel of Figure 2 are cells that have internalized PEG-CNARHO. In contrast, the lack of fluorescence in the other panels (PEG-RHO and RHO alone) indicates that comparatively less uptake was observed. Semiquantitative analysis confirms that cells incubated with PEG-CNA-RHO exhibited a statistically higher average fluorescence intensity than cells treated with C

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

Subcellular localization of the CNA copolymers was studied through immunofluorescence microscopy. In short, cells were treated with a known concentration of PEG-CNA-RHO (red fluorescence) for 1 h, fixed, and then stained with organellespecific fluorophore−antibody conjugates (green fluorescence). By measuring the presence and intensity of red fluorescence against green fluorescence for each pixel of an image, a semiquantitative degree of colocalization is determined in the form of the Pearson coefficient or Mander’s coefficient.29,30 This phenomenon is also visually identified in an overlay image by the presence of yellow pixels. In all, four organelles were stained: endosomes, endoplasmic reticulum, mitochondria, and lysosomes. Additionally, nuclei and actin were stained to aid in cellular identification. Resulting overlay images of individual cells are shown in Figure 4. The

Figure 2. Cellular internalization of PEG-CNA-RHO, PEG-RHO, and RHO. Cells were treated with 100 μg/mL of conjugate or an equimolar amount of rhodamine for 1 h, washed, fixed, and imaged by fluorescence microscopy. A significantly greater average fluorescence intensity per cell after washing was determined for cells treated with PEG-CNA-RHO (*p < 0.001) than for cells treated with PEG-RHO or RHO controls, indicating CNA dependent cellular uptake. (n > 1000, error bars represent standard deviations, scale bar = 500 μm.)

PEG-RHO or RHO alone. These results suggest that the presence of CNA in the copolymer facilitates uptake by cells. Uptake kinetics were investigated by incubating cells with 100 μg/mL of PEG-CNA-RHO for various times ranging from 30 s to 1 h. This concentration was chosen because it was shown to be cytocompatible as well as visible with microscopy. Representative confocal microscopy images and semiquantitative image analysis show an increase in cell-associated fluorescence over time, which are shown in Figure 3. All further experiments were conducted with a 1 h incubation step because it appears that PEG-CNA-RHO uptake levels are easily visualized on this time scale.

Figure 4. Three representative images of immunofluorescence microscopy images highlight the cellular localization of PEG-CNArho conjugates. Cells were treated with 100 μg/mL of PEG-CNA-rho conjugates for 1 h, then fixed, permeabilized, and stained. Blue represents the nucleus (Hoechst 33342), red represents the PEGCNA-RHO conjugate, and green represents either endosomes, the endoplasmic reticulum, mitochondria, or lysosomes. Areas of yellow indicate colocalization between the PEG-CNA-rho conjugate and the indicated organelle. Areas of magenta indicate colocalization between the PEG-CNA-rho conjugate and the nucleus. (scale bar = 10 μm.)

localization of PEG-CNA-RHO (red fluorescence) in each image shows a similar motif: the presence of small punctate areas of high fluorescence amid a mostly diffuse pattern within the cell. The small punctate areas within these images could suggest the sequestration of PEG-CNA-RHO by an organelle or could simply be aggregation of the polymer in the highly complex intracellular environment. One phenomenon to mention is that in some images it appears that the PEGCNA-RHO and organelles look to be negatively correlated. This behavior is seen in the images staining for mitochondria, lysosomes, and the endoplasmic reticulum and are indicated by the white arrows. It should also be noted that there is evidence of PEG-CNA-RHO in the nucleus, which is seen by the magenta color in the images in Figure 4. Previous studies have

Figure 3. Incubating cells for increasingly longer times leads to an increased average cell-associated fluorescence. Representative images show cell fluorescence at 30 s, 10 min, 30 min, and 1 h. Fluorescent intensity appears to begin to level off on the order of hours. (n > 30, error bars represent standard deviations, scale bar = 25 μm.) D

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules

philic molecules have indeed shown that increasing hydrophobicity leads to greater cell uptake.36 Further, studies on amphiphilic cell-penetrating peptides (CPPs) have concluded that insertion of the CPPs hydrophobic region into the cell membrane is crucial for the initial steps of membrane translocation.37,38 It is also worth noting that like many other amphiphilic molecules, these polymers self-assemble into particles or aggregates in solution (Figure S10). There have been other reports of self-assembled structures that are internalized by energy-independent pathways; however, the exact mechanisms are not fully understood.39,40 In addition, CNAs maintain a net neutral charge under biological conditions, which could also affect their interaction with the cell membrane. It has been shown that electrostatics affect the hydrophobic partitioning of amphiphiles into lipid-bilayer membranes.38,41 This behavior suggests that neutral CNAs have an increased association with the cell membrane and ultimately increased uptake compared to negatively charged DNA, which experiences electrostatic repulsion between the negatively charged backbone of DNA and the negatively charged cell membrane surface.42,43

shown that nuclear uptake of PS-DNA (and to some extent, natural phosphodiester linked DNA) occurs to an extent.31,32 However, some reports have warned that this result could be an artifact of the fixation and permeabilization process.33 To calculate the Pearson and Mander’s coefficients, the JACoP ImageJ plugin was used to determine the overlap between the green and red channels of the overlay images. The Pearson coefficient generally measures how well the intensities of each channel are correlated for every pixel of an image. The Mander’s coefficient takes this one step further and can differentiate between overlap of the green channel with the red or the red channel with the green. For both coefficients, only values close to 1 indicate significant colocalization, while all other values do not allow conclusions to be drawn.34 In all images, colocalization coefficients were no higher than 0.4, suggesting that no conclusions can be made regarding colocalization between PEG-CNA-RHO and any of the organelles examined. However, this does not mean that there is no colocalization between PEG-CNA-RHO and subcellular organelles. It is possible that significant interactions are occurring, but the uncertainty associated with immunofluorescence techniques does not allow for its detection. For instance, given that the CNA block contains repeats of thymine, we might expect the copolymer to bind to structures containing adenine residues, like the poly-A tail of mRNA. However, the immunofluorescence techniques here do not allow for the detection of such specific interactions. To identify the mechanism by which uptake of CNA occurs, experiments were performed at 4 or 37 °C to probe passive and active transport mechanisms, respectively. At 4 °C, active endocytotic uptake mechanisms are severely limited due to the lack of energy (i.e., ATP production),35 but passive transport pathways are not. Thus, the relative extent of uptake that is observed at 4 °C provides information about whether uptake is passive or active. Cells were incubated at either 4 or 37 °C and then treated with PEG-CNA-RHO. As seen in Figure 5, the



CONCLUSIONS CNAs have been developed as an alternative approach to modifying nucleic acids. Their ability to participate in thiol−X reactions allows for facile synthesis of nucleic acid oligomers. Similar oligomers have been shown to have effective antisense characteristics in gene therapy applications. In the present study, the evaluation of CNA-based copolymers as potential antisense oligonucleotides was initiated by evaluating their cytocompatibility and cellular internalization characteristics. CNA copolymers were determined to be nontoxic to cells at concentrations up to 100 μg/mL. Additionally, PEG-CNARHO was taken up by cells within 1 h and to a greater extent than PEG-RHO and RHO alone, and the extent of uptake was not affected by temperature, suggesting a passive, CNAdependent uptake mechanism. It is possible that this uptake is because the lipophilic CNA block allows for better interaction with the lipid cell membrane. The distribution of PEG-CNA-RHO within cells exhibited both diffuse and punctate patterns with no apparent specific organelle colocalization. These results show that CNAs are cytocompatible, readily taken up by cells, and thus warrant future studies to assess their potential as a synthetically simple alternative to current antisense approaches.



Figure 5. Temperature-dependent uptake of PEG-CNA-RHO. Cells were incubated with 100 μg/mL of PEG-CNA-RHO for 1 h at either 4 or 37 °C. Cellular uptake of PEG-CNA-RHO was not significantly affected by incubation temperature, suggesting a passive uptake mechanism. (n > 45, error bars represent standard deviations, scale bar = 100 μm.)

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.biomac.8b00162. 1 H NMR of intermediates for PEG-CNA-RHO and PEG-RHO synthesis (PEG-SH, thymine thiol−ene monomer, PEG-CNA, PEG-CNA-RHO, PEG-RHO); GPC traces of intermediates; UV−vis of PEG-CNARHO and method of calculating extent of labeling by fluorescent dye; DLS data showing size distribution of particles formed (PDF)

difference in average cellular fluorescence was determined to be insignificant between 4 and 37 °C, suggesting that uptake is primarily through a passive mechanism (i.e., free or facilitated diffusion). It is likely that some of these results are explained by unique structural elements of the CNAs. First, the PEG-CNA-RHO exhibits block amphiphilicity due to the hydrophilic PEG block and hydrophobic CNA block. It is possible that the hydrophobic CNA block is facilitating interaction with the phospholipids within the cell membrane, leading to increased translocation. Previous reports of similarly structured amphi-



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (S.J.B.). E

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules ORCID

(11) Wickstrom, E. Oligodeoxynucleotide Stability in Subcellular Extracts and Culture Media. J. Biochem. Biophys. Methods 1986, 13 (2), 97−102. (12) Grunweller, A. Comparison of Different Antisense Strategies in Mammalian Cells Using Locked Nucleic Acids, 2’-O-Methyl RNA, Phosphorothioates and Small Interfering RNA. Nucleic Acids Res. 2003, 31 (12), 3185−3193. (13) Sazani, P.; Kang, S.-H.; Maier, M. A.; Wei, C.; Dillman, J.; Summerton, J.; Manoharan, M.; Kole, R. Nuclear Antisense Effects of Neutral, Anionic and Cationic Oligonucleotide Analogs. Nucleic Acids Res. 2001, 29 (19), 3965−3974. (14) Dueholm, K. L.; Egholm, M.; Behrens, C.; Christensen, L.; Hansen, H. F.; Vulpius, T.; Petersen, K. H.; Berg, R. H.; Nielsen, P. E.; Buchardt, O. Synthesis of Peptide Nucleic Acid Monomers Containing the Four Natural Nucleobases: Thymine, Cytosine, Adenine, and Guanine and Their Oligomerization. J. Org. Chem. 1994, 59 (19), 5767−5773. (15) Koshkin, A. A.; Singh, S. K.; Nielsen, P.; Rajwanshi, V. K.; Kumar, R.; Meldgaard, M.; Olsen, C. E.; Wengel, J. LNA (Locked Nucleic Acids): Synthesis of the Adenine, Cytosine, Guanine, 5Methylcytosine, Thymine and Uracil Bicyclonucleoside Monomers, Oligomerisation, and Unprecedented Nucleic Acid Recognition. Tetrahedron 1998, 54 (14), 3607−3630. (16) Badi, N.; Lutz, J.-F. Sequence Control in Polymer Synthesis. Chem. Soc. Rev. 2009, 38 (12), 3383. (17) Xi, W.; Pattanayak, S.; Wang, C.; Fairbanks, B.; Gong, T.; Wagner, J.; Kloxin, C. J.; Bowman, C. N. Clickable Nucleic Acids: Sequence-Controlled Periodic Copolymer/Oligomer Synthesis by Orthogonal Thiol-X Reactions. Angew. Chem., Int. Ed. 2015, 54 (48), 14462−14467. (18) Spitzer, S.; Eckstein, F. Inhibition of Deoxyribonucleases by Phosphorothioate Groups in Oligodeoxyribonucleotides. Nucleic Acids Res. 1988, 16 (24), 11691−11704. (19) Crinelli, R.; Bianchi, M.; Gentilini, L.; Magnani, M. Design and Characterization of Decoy Oligonucleotides Containing Locked Nucleic Acids. Nucleic Acids Res. 2002, 30 (11), 2435−2443. (20) Aronin, N.; DiFiglia, M. Huntingtin-Lowering Strategies in Huntington’s Disease: Antisense Oligonucleotides, Small RNAs, and Gene Editing: HUNTINGTIN-LOWERING STRATEGIES IN HD. Mov. Disord. 2014, 29 (11), 1455−1461. (21) Harguindey, A.; Domaille, D. W.; Fairbanks, B. D.; Wagner, J.; Bowman, C. N.; Cha, J. N. Synthesis and Assembly of Click-NucleicAcid-Containing PEG−PLGA Nanoparticles for DNA Delivery. Adv. Mater. 2017, 29 (24), 1700743. (22) Harris, J. M.; Martin, N. E.; Modi, M. Pegylation: A Novel Process for Modifying Pharmacokinetics. Clin. Pharmacokinet. 2001, 40 (7), 539−551. (23) Harada, A.; Togawa, H.; Kataoka, K. Physicochemical Properties and Nuclease Resistance of Antisense-Oligodeoxynucleotides Entrapped in the Core of Polyion Complex Micelles Composed of Poly (Ethylene Glycol)−poly (L-Lysine) Block Copolymers. Eur. J. Pharm. Sci. 2001, 13 (1), 35−42. (24) Jia, F.; Lu, X.; Tan, X.; Wang, D.; Cao, X.; Zhang, K. Effect of PEG Architecture on the Hybridization Thermodynamics and Protein Accessibility of PEGylated Oligonucleotides. Angew. Chem., Int. Ed. 2017, 56 (5), 1239−1243. (25) Chan, W. S.; Svensen, R.; Phillips, D.; Hart, I. R. Cell Uptake, Distribution and Response to Aluminium Chloro Sulphonated Phthalocyanine, a Potential Anti-Tumour Photosensitizer. Br. J. Cancer 1986, 53 (2), 255−263. (26) Bryant, S. J.; Nuttelman, C. R.; Anseth, K. S. Cytocompatibility of UV and Visible Light Photoinitiating Systems on Cultured NIH/ 3T3 Fibroblasts in Vitro. J. Biomater. Sci., Polym. Ed. 2000, 11 (5), 439−457. (27) Jin, H.; Heller, D. A.; Strano, M. S. Single-Particle Tracking of Endocytosis and Exocytosis of Single-Walled Carbon Nanotubes in NIH-3T3 Cells. Nano Lett. 2008, 8 (6), 1577−1585. (28) Lönnroth, E.-C. Toxicity of Medical Glove Materials: A Pilot Study. Int. J. Occup. Saf. Ergon. 2005, 11 (2), 131−139.

Christopher N. Bowman: 0000-0001-8458-7723 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was completed with support from an NSF MRSEC grant (DMR 1420736) and from a US Department of Education GAANN Fellowship to Alex Anderson. E.B.P. was supported by the NIH T32 National Institutional Research Service Award T32 HL07670. The imaging work was performed at the BioFrontiers Institute Advanced Light Microscopy Core. Spinning disc confocal microscopy was performed on Nikon Ti-E microscope supported by the BioFrontiers Institute and the Howard Hughes Medical Institute.



ABBREVIATIONS CNA, “clickable” nucleic acid; PEG, poly(ethylene glycol); PEG-CNA-RHO, fluorescent clickable nucleic acid copolymer; PEG-RHO, fluorescent poly(ethylene glycol) polymer; RHO, rhodamine.



REFERENCES

(1) Olie, R. A.; Simões-Wüst, A. P.; Baumann, B.; Leech, S. H.; Fabbro, D.; Stahel, R. A.; Zangemeister-Wittke, U. A Novel Antisense Oligonucleotide Targeting Survivin Expression Induces Apoptosis and Sensitizes Lung Cancer Cells to Chemotherapy. Cancer Res. 2000, 60 (11), 2805−2809. (2) Savage, D. B.; Choi, C. S.; Samuel, V. T.; Liu, Z.-X.; Zhang, D.; Wang, A.; Zhang, X.-M.; Cline, G. W.; Yu, X. X.; Geisler, J. G.; Bhanot, S.; Monia, B. P.; Shulman, G. I. Reversal of Diet-Induced Hepatic Steatosis and Hepatic Insulin Resistance by Antisense Oligonucleotide Inhibitors of Acetyl-CoA Carboxylases 1 and 2. J. Clin. Invest. 2006, 116 (3), 817−824. (3) Zinker, B. A.; Rondinone, C. M.; Trevillyan, J. M.; Gum, R. J.; Clampit, J. E.; Waring, J. F.; Xie, N.; Wilcox, D.; Jacobson, P.; Frost, L. PTP1B Antisense Oligonucleotide Lowers PTP1B Protein, Normalizes Blood Glucose, and Improves Insulin Sensitivity in Diabetic Mice. Proc. Natl. Acad. Sci. U. S. A. 2002, 99 (17), 11357−11362. (4) Paterson, B. M.; Roberts, B. E.; Kuff, E. L. Structural Gene Identification and Mapping by DNA-mRNA Hybrid-Arrested CellFree Translation. Proc. Natl. Acad. Sci. U. S. A. 1977, 74 (10), 4370− 4374. (5) Zamecnik, P. C.; Stephenson, M. L. Inhibition of Rous Sarcoma Virus Replication and Cell Transformation by a Specific Oligodeoxynucleotide. Proc. Natl. Acad. Sci. U. S. A. 1978, 75 (1), 280−284. (6) Shi, B.; Zheng, M.; Tao, W.; Chung, R.; Jin, D.; Ghaffari, D.; Farokhzad, O. C. Challenges in DNA Delivery and Recent Advances in a Multifunctional Polymeric DNA Delivery Systems. Biomacromolecules 2017, 18, 2231. (7) Kafil, V.; Omidi, Y. Cytotoxic Impacts of Linear and Branched Polyethylenimine Nanostructures in A431 Cells. BioImpacts BI 2011, 1 (1), 23−30. (8) Breunig, M.; Lungwitz, U.; Liebl, R.; Goepferich, A. Breaking up the Correlation between Efficacy and Toxicity for Nonviral Gene Delivery. Proc. Natl. Acad. Sci. U. S. A. 2007, 104 (36), 14454−14459. (9) Lechardeur, D.; Sohn, K.-J.; Haardt, M.; Joshi, P. B.; Monck, M.; Graham, R. W.; Beatty, B.; Squire, J.; O’Brodovich, H.; Lukacs, G. L. Metabolic Instability of Plasmid DNA in the Cytosol: A Potential Barrier to Gene Transfer. Gene Ther. 1999, 6 (4), 482. (10) Eder, P. S.; DeVINE, R. J.; Dagle, J. M.; Walder, J. A. Substrate Specificity and Kinetics of Degradation of Antisense Oligonucleotides by a 3′ Exonuclease in Plasma. Antisense Res. Dev. 1991, 1 (2), 141− 151. F

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX

Article

Biomacromolecules (29) Zinchuk, V.; Grossenbacher-Zinchuk, O. Quantitative Colocalization Analysis of Confocal Fluorescence Microscopy Images. In Current Protocols in Cell Biology; Bonifacino, J. S., Dasso, M., Harford, J. B., Lippincott-Schwartz, J., Yamada, K. M., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, 2011. (30) Manders, E. M. M.; Verbeek, F. J.; Aten, J. A. Measurement of Co-Localization of Objects in Dual-Colour Confocal Images. J. Microsc. 1993, 169 (3), 375−382. (31) Lorenz, P.; Baker, B. F.; Bennett, C. F.; Spector, D. L. Phosphorothioate Antisense Oligonucleotides Induce the Formation of Nuclear Bodies. Mol. Biol. Cell 1998, 9 (5), 1007−1023. (32) Leonetti, J. P.; Mechti, N.; Degols, G.; Gagnor, C.; Lebleu, B. Intracellular Distribution of Microinjected Antisense Oligonucleotides. Proc. Natl. Acad. Sci. U. S. A. 1991, 88 (7), 2702−2706. (33) Melan, M. A.; Sluder, G. Redistribution and Differential Extraction of Soluble Proteins in Permeabilized Cultured Cells. Implications for Immunofluorescence Microscopy. J. Cell Sci. 1992, 101 (Pt 4), 731−743. (34) Bolte, S.; Cordelieres, F. P. A Guided Tour into Subcellular Colocalization Analysis in Light Microscopy. J. Microsc. 2006, 224 (3), 213−232. (35) Schmid, S. L.; Carter, L. L. ATP Is Required for ReceptorMediated Endocytosis in Intact Cells. J. Cell Biol. 1990, 111 (6), 2307−2318. (36) Luxenhofer, R.; Sahay, G.; Schulz, A.; Alakhova, D.; Bronich, T. K.; Jordan, R.; Kabanov, A. V. Structure-Property Relationship in Cytotoxicity and Cell Uptake of poly(2-Oxazoline) Amphiphiles. J. Controlled Release 2011, 153 (1), 73−82. (37) Deshayes, S.; Heitz, A.; Morris, M. C.; Charnet, P.; Divita, G.; Heitz, F. Insight into the Mechanism of Internalization of the CellPenetrating Carrier Peptide Pep-1 through Conformational Analysis. Biochemistry 2004, 43 (6), 1449−1457. (38) Seelig, J. Thermodynamics of Lipid−peptide Interactions. Biochim. Biophys. Acta, Biomembr. 2004, 1666 (1−2), 40−50. (39) Pujals, S.; Fernández-Carneado, J.; López-Iglesias, C.; Kogan, M. J.; Giralt, E. Mechanistic Aspects of CPP-Mediated Intracellular Drug Delivery: Relevance of CPP Self-Assembly. Biochim. Biophys. Acta, Biomembr. 2006, 1758 (3), 264−279. (40) Pujals, S.; Giralt, E. Proline-Rich, Amphipathic Cell-Penetrating Peptides. Adv. Drug Delivery Rev. 2008, 60 (4), 473−484. (41) Li, Y.; Heitz, F.; Le Grimellec, C.; Cole, R. B. Fusion Peptide− Phospholipid Noncovalent Interactions As Observed by Nanoelectrospray FTICR−MS. Anal. Chem. 2005, 77 (6), 1556−1565. (42) Patil, S. D.; Rhodes, D. G.; Burgess, D. J. DNA-Based Therapeutics and DNA Delivery Systems: A Comprehensive Review. AAPS J. 2005, 7 (1), E61−E77. (43) Meade, B. R.; Dowdy, S. F. Exogenous siRNA Delivery Using Peptide Transduction Domains/Cell Penetrating Peptides. Adv. Drug Delivery Rev. 2007, 59 (2), 134−140.

G

DOI: 10.1021/acs.biomac.8b00162 Biomacromolecules XXXX, XXX, XXX−XXX