129
Graduate eses and Dissertations Iowa State University Capstones, eses and Dissertations 2017 Selection of functional RNA aptamers against Ebola glycoproteins Shambhavi Shubham Iowa State University Follow this and additional works at: hps://lib.dr.iastate.edu/etd Part of the Biochemistry Commons is Dissertation is brought to you for free and open access by the Iowa State University Capstones, eses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Graduate eses and Dissertations by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected]. Recommended Citation Shubham, Shambhavi, "Selection of functional RNA aptamers against Ebola glycoproteins" (2017). Graduate eses and Dissertations. 16528. hps://lib.dr.iastate.edu/etd/16528

Selection of functional RNA aptamers against Ebola

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Selection of functional RNA aptamers against Ebola

Graduate Theses and Dissertations Iowa State University Capstones, Theses andDissertations

2017

Selection of functional RNA aptamers againstEbola glycoproteinsShambhavi ShubhamIowa State University

Follow this and additional works at: https://lib.dr.iastate.edu/etd

Part of the Biochemistry Commons

This Dissertation is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State UniversityDigital Repository. It has been accepted for inclusion in Graduate Theses and Dissertations by an authorized administrator of Iowa State UniversityDigital Repository. For more information, please contact [email protected].

Recommended CitationShubham, Shambhavi, "Selection of functional RNA aptamers against Ebola glycoproteins" (2017). Graduate Theses and Dissertations.16528.https://lib.dr.iastate.edu/etd/16528

Page 2: Selection of functional RNA aptamers against Ebola

Selection of functional RNA aptamers against Ebola glycoproteins

by

Shambhavi Shubham

A dissertation submitted to the graduate faculty

in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY

Major: Molecular Cellular and Developmental Biology

Program of Study Committee

Marit Nilsen-Hamilton, Major Professor

W. Allen Miller

Eric Henderson

Drena Dobbs

Walter Moss

Iowa State University

Ames, Iowa

2017

Copyright © Shambhavi Shubham 2017. All rights reserved.

Page 3: Selection of functional RNA aptamers against Ebola

ii

DEDICATION

I dedicate this dissertation to my mother Saroj Shrivastava for her constant support and

inspiration.

Page 4: Selection of functional RNA aptamers against Ebola

iii

TABLE OF CONTENTS

Page

LIST OF FIGURES…………………………………………………………………...v

ACKNOWLEDGMENTS……………………………………………………………vii

ABSTRACT……………………………………………………………………….....viii

CHAPTER 1: REVIEW OF LITERATURE

Filoviruses……………………………………………………………………..1

Filoviruses Pathogenesis………………………………………………………2

Ebola Virus Immune Evasion Strategy………………………………………..3

Viral Replication And Transcription…………………………………………..5

Ebola Glycoprotein GP1, 2……………………………………………………7

Ebola Glycoprotein GP1, 2 And Its Interaction with the Host Cell…………...9

Ebola Glycoprotein GP1, 2 Mediating Host and Cell Membrane Fusion……11

Ebola sGP…………………………………………………………………….12

Ebola sGP Role In Pathogenesis……………………………………………..14

Ebola Therapeutics And Vaccine…………………………………………….15

Ebola Diagnostic Kits………………………………………………………...16

Aptamers……………………………………………………………………...17

Methods of Selection In Aptamers……………………………………………21

High Throughput Sequencing And Data Analysis……………………………25

Aptamers vs Antibodies And Its Market State………………………………..27

Page 5: Selection of functional RNA aptamers against Ebola

iv

Aptamers In Diagnosis Of Viral Infections…………………………………...29

Aptamers In Preventing The Fusion Of Virus Particle With Host Cell……….31

CHAPTER 2: A 2’FY RNA EPITOPE FORMS AN APTAMER FOR

EBOLAVIRUS SGP: SELECTION AND CHARACTERIZATION………………..60

Abstract………………………………………………………………………..61

Introduction………………………………………………………………........62

Results……………………………………………………………………........63

Discussion……………………………………………………………………..67

Materials And Methods……………………………………………………….71

Legends To Figures……………………………………………………….......77

References……………………………………………………………………..79

CHAPTER 3: SELECTING A FUNCTIONAL ANTI–VIRAL RNA APTAMER

AGAINST EBOLAVIRUS SURFACE GLYCOPROTEIN…………………………..88

Abstract………………………………………………………………………...89

Introduction…………………………………………………………………......90

Results………………………………………………………………………......92

Discussion………………………………………………………………………95

Materials and Methods………………………………………………………….99

Legends to Figures…………………………………………………………......104

References……………………………………………………………………...106

CHAPTER4: CONCLUSION AND DISCUSSION………………………………...115

References………………………………………………………………………121

Page 6: Selection of functional RNA aptamers against Ebola

v

LIST OF FIGURES

CHAPTER 1 Page

Figure 1 Ebola Virus Particle 34

Figure 2 GP1,2Δmucin X ray Crystallography Structure 35

Figure 3 Prefusion and postfusion GP1,2 conformation 36

Figure 4 Structures of full length GP1,2 and sGP 37

Figure 5 SELEX protocol to isolate RNA aptamers 38

CHAPTER 2

Figure 1 Next Gen Sequencing Analysis and Consensus Motif Identification 83

Figure 2 Determining the affinities of 2’FY oligonucleotides against sGP 84

Figure 3 Determining the binding epitope of sGP on 2’FY RNA aptamer 85

Figure 4 Competition Binding Assay to determine the site of interaction of 86

various aptamer sequences

Figure 5 Specificity of 2’FY aptamer for sGP over serum proteins 87

CHAPTER 3

Figure 1 Comparison of top 10 sequences from SELEX experiments to select

aptamers that recognize GP1,2Δmucin and full length GP1,2 111

Figure 2 Selection through a sucrose cushion 112

Figure 3 Filter capture assay to estimate the affinity of oligonucleotide 4789 113

for GP1,2Δmucin

Page 7: Selection of functional RNA aptamers against Ebola

vi

Figure 4 Filter capture assay to determine the affinity of oligonucleotide 5185 114

and 5187 for GP1,2

Page 8: Selection of functional RNA aptamers against Ebola

vii

ACKNOWLEDGMENTS

I would like to thank my advisor Dr. Marit Nilsen-Hamilton for her support and

guidance all these years. She has taught me how to be a good experimentalist and an

independent researcher. Her ideas and knowledge made my Ph.D. experience productive and

stimulating. I am thankful to my committee members Dr. Allen Miller, Dr. Drena Dobbs, Dr.

Eric Henderson and Dr. Walter Moss for their time, guidance, and insightful questions. I would

like to acknowledge Dr. Eric Underbakke for teaching me Mass Spectrometry and being a

great mentor.

I would like to thank my lab manager Lee Bendickson for being very patient with me

and teaching critical technical details essential for designing good experiments. Thank you to

my past and present lab members who have contributed immensely to my personal and

professional time at Iowa State University. They have provided me with great guidance and

collaboration all these years and have been great friends as well.

My time in Ames was made enjoyable due to my friends, who provided me with a lot

of love, support, and guidance. I am grateful for the time spent with them not just discussing

science but also life.

Lastly, I would like to thank my parents and my sister for their love and encouragement.

This journey wouldn’t have been possible without them. Thanks to my loving, supporting and

patient husband Gunjan Pandey for standing by me through thick and thin. And most of all my

very loving son Reyansh Pandey, he’s made my life complete and encourages me to be a better

person every day. Thank You.

Page 9: Selection of functional RNA aptamers against Ebola

viii

ABSTRACT

Ebola viruses (EBOV) cause severe disease symptoms in humans and in non-human

primates in the form of viral hemorrhagic fever. Although EBOV outbreaks have not occurred

in the U.S., the virus is of public health concern as a potential bioterrorism organism for which

no vaccine or anti-viral is available. Fast-acting and prophylactic therapeutics are needed to

reduce mortality due to outbreaks in other countries and for use in the U.S. if the virus is used

in a bioterrorism attack. In view of the paucity of current antiviral therapies and diagnostic

systems for EBOV, we provide an alternative solution by selection of high affinity 2’FY-

stabilized RNA aptamers that bind the EBOV surface exposed glycoprotein, GP1 and soluble

glycoprotein (sGP). Aptamers are single stranded short nucleic acid oligonucleotides with

sequences that enable high affinity and specificity for their targets. Aptamers have comparable

affinity with antibodies, but they are not immunogenic and are raised by in-vitro methods.

They can be selected to bind to a precise region of a protein. By this means, an aptamer binding

the EBOV surface GP1 would prevent the interaction between the virus and the host cell,

disrupt the viral life cycle and an aptamer against sGP can be integrated with a detection

platform that can be used as a biosensor to detect Ebola infections.

Page 10: Selection of functional RNA aptamers against Ebola

1

CHAPTER 1

INTRODUCTION

Filovirus Overview

Filoviridae family is composed of three genera: Ebolaviruses, Marburgviruses and

Cueva viruses. There are two species of Marburgvirus, Marburg virus (MARV) and Ravn

virus (RAVV) and five species of Ebolavirus, Ebola virus (EBOV), Sudan virus (SUDV), Tai

Forest virus (TAFV), Reston Virus (RESTV) and Bundibugyo virus (BDBV) [1]. Ebola virus

is an enveloped, non-segmented and negatively stranded RNA virus with genome of roughly

19kb. Ebola virus particle is filamentous in shape and genome encodes seven structural

proteins beginning from the 3’ end: the nucleoprotein (NP), viral protein (VP35), VP40, the

glycoprotein (GP), VP30, VP24 and the RNA dependent RNA polymerase (L). The viral RNA

genome is encapsulated in NP that, along with VP35, VP30, and L form the replication

complex [2] (Figure 1). VP 40 is the major matrix protein and is sufficient for budding of virus-

like particles (VLPs), VP24 is a minor matrix protein [3, 4]. In addition to these proteins, Ebola

virus secretes into the blood stream soluble glycoprotein (sGP). It is a nonstructural, secretory

glycoprotein, which shares a 295-amino acid sequence with the glycoprotein GP. Ebolavirus

infection can cause hemorrhagic fever with mortality rates as high as 90% [5]. The most recent

outbreak was in 2014 when the incidence and prevalence of Ebola infections exceeded any

previous outbreak [6]. With its ability to cause severe pathogenicity and a high mortality rate,

the potential of being used for bioterrorism and the absence of any licensed vaccines or

therapeutics underscores the threat, this virus poses to public health [7].

Page 11: Selection of functional RNA aptamers against Ebola

2

Filovirus Pathogenesis

Filoviruses cause hemorrhagic fever in human and non-human primates and exhibit

high rates of mortality reaching up to 90%. Initial infection occurs through direct transmission

from bats which are believed to be the host reservoir [4]. Ebola virus enters the host through

mucosal surfaces and breaks through the skin. Most human infections have been caused by

direct contact with the human patients or cadavers or infected animal carriers. Infectious virus

particles or viral RNA have been detected in the semen and genital secretions of infected

patients [4]. The known infectivity titers in non-human primates have been reported to be in

the range of 107 to 108 plaque forming units/g[8], any exposure though the oral route can be

associated with very high infectious doses. Transmission of the Ebola virus through an aerosol

route has not been documented and is thought to be rare. Ebola virus, once it gains entry into

the body, targets a range of cells early during infection, including monocytes, macrophages

and dendritic cells, and then spreads to a variety of other cell types, including epithelial cells

in the visceral organs [10]. Transport of infected immune cells and free virus through the blood

stream is believed to enable the virus to spread from the site of initial infection. The high

lethality of Ebola viruses in large part is due to an early and strong inhibition of the innate

immune response [20].

Page 12: Selection of functional RNA aptamers against Ebola

3

Ebola Virus Immune Evasion Strategies

Phosphatidyl serine dependent

Ebola virus particles are filamentous in shape and can be up to a micron in length,

which would make it difficult for them to be taken in by the cells by clathrin or caveolin

mediated endocytosis. Clathrin mediated endocytosis is usually limited to virions of 100nm

whereas caveolin is around 50nm. It could be for this reason that Ebola virus is taken up by

macropinocytosis. Phosphatidylserine is a lipid found on the outer leaflet of apoptotic cells

where it alerts phagocytic cells to engulf the apoptotic cells by macropinocytosis. In a similar

fashion, phosphatidylserine on the Ebola virus particles is recognized by the (T-cell

immunoglobulin and mucin domain 1) TIM-1 receptors, interaction between TIM-1 receptor

and phosphatidylserine mediates the entry of virus into the host cells. Ebola virus acquires

phosphatidylserine from the plasma membrane microdomains, called lipid rafts, during the

budding process. Lipid rafts are highly enriched for phosphatidylserine on the external

leaflet[11]. The Ebola virus particles present themselves to phagocytic cells as apoptotic

bodies and avoid initiating the inflammatory response by being engulfed by the cells. [12, 13].

Page 13: Selection of functional RNA aptamers against Ebola

4

Preventing the interferon response

VP35

VP35 is a multifunctional protein, which is associated with the ribonuclear complex

and links the viral RNA polymerase (L) to the nucleoprotein (NP) during viral replication.

VP35 also inhibits α and β interferon responses by multiple mechanisms. One mechanism is

mediated by inhibiting RIG-I and MDA-5, innate pattern recognition receptors that detect

foreign cytoplasmic RNA. RIG-I recognizes 5’-triphoshophates of the blunt end of dsRNA and

MDA-5 senses long dsRNA. RIG-I and MDA-5 associates with a common downstream

adaptor IPS-1, which activates NF-kB, IRF-3, and IRF- 7 to trigger the expression of type I

interferon and pro-inflammatory cytokines. VP35 is a decoy substrate for IKKε /TBK-1

kinases and impedes the RIG-1 pathway by binding the N-terminal kinase domain of IKKε and

prevents IRF-3 phosphorylation. Interactions of IKKε with other proteins such as IRF-1 and

IPS-1 are also prevented due to binding of VP35 to IKKε [14]. These interactions successfully

block the activation of genes with interferon responsive promoters [15, 16]. In a second

mechanism, VP35 competes with RIG-I by binding to the blunt end of dsRNA, which prevents

RIG-I and MDA-5 responses [17-19].

VP24

Typically, when innate immunity is functional, the host response to a virus infection is

secretion of interferon proteins to generate antiviral responses in neighboring cells, signal

hematopoietic cell responses and increase antigen presentation. Secreted interferon binds to

type I and II interferon receptors, activating signaling via adaptor proteins, which results in the

Page 14: Selection of functional RNA aptamers against Ebola

5

phosphorylation and dimerization of the signal transducer and activator of transcription

(STAT) proteins. These dimerized transcription factors are transported to the nucleus where

they bind to interferon response elements and induce antiviral gene expression [20]. Since,

these pathways are so important in fighting viral infections, viruses usually target them.

Karyopherin-α1 is the nuclear membrane localized signal receptor that is responsible for

nuclear accumulation of the STAT-1 dimers. VP24 interacts with karyopherin-α1 to inhibit the

nuclear accumulation of STAT-1 and block interferon signaling. This mechanism prevents

antiviral effects in cells expressing Ebola virus [21].

RNA stability and replication

The heterogeneous nuclear protein complexes, C1/C2 proteins ((hnRNP C1/C2), reside

in the nucleus where they bind poly-U mRNA and assist in splicing before they are transported

into the cytoplasm. They also engage in cap independent, IRES (Internal Ribosome Entry Site)

dependent translation in the cytoplasm during mitosis [22]. Ebola VP24 causes relocalization

of hnRNP C1/C2 from the nucleus to cytoplasm. It is hypothesized that C1/C2 might bind the

poly-U tracts in Ebola viral RNA, thereby stabilizing it against degradation and enhancing

genomic replication [22]. This suggests that Ebola virus uses the host machinery to prolong

the half-life of its viral RNA and optimize transcription from its genome and viral replication.

Filoviral Replication and Transcription

Filovirus replication and transcription happens in the cytoplasm of the infected cell, the

virus encodes its own RNA-dependent RNA polymerase that uses the negative-strand viral

Page 15: Selection of functional RNA aptamers against Ebola

6

RNA as a template for transcription and replication. Nucleocapsid assembly for the virus is

mediated by three proteins NP, VP35 and VP30. The rate-limiting step for nucleocapsid

formation is nucleoprotein (NP) assisted formation of helical bundle. [23]. This is followed by

interactions with VP35 and VP30 and L to produce the ribonucleoprotein that encapsulates the

viral genome. [24, 25]. For budding and formation of a virion, VP35 interacts with capsid

protein VP40 and mediates the packaging of the nucleocapsid into the mature virion on the cell

surface [26].

Replication of the RNA genome is initiated by the synthesis of positive sense

intermediate (antigenome). The antigenome is the reverse complement of the RNA genome

and it serves as a template for the generation of negative stranded RNA. Studies with Marburg

and Ebola virus minigenome systems have established NP, VP35 and L as the minimal

essential components for viral replication [27, 28] [29].

Transcription process begins after the viral RNA is released into cytoplasm, the

polymerase complex starts to transcribe it. Viral genome available for transcription is oriented

in 3’ to 5’ direction, due to the polymerase complex stopping and reinitiating at each gene

junction, transcription of the starting gene NP is favored. Due to sequential transcription, a

gradient of transcribed gene products is generated. This leads to a non-homogenous

distribution of viral proteins [30]. During transcription the viral RNA is polyadenylated as

well as capped for cap-dependent translation of the viral mRNA [31].

Little is known about the mechanism of switching from the transcription to the

replication state but its hypothesized that the phosphorylation state of VP30 might play a

Page 16: Selection of functional RNA aptamers against Ebola

7

critical role in deciding if the polymerase complex will support replication or transcription

[32].

Viral assembly and budding are mediated by VP24 and VP40. To prepare the virus for

budding VP24, one of the matrix proteins, associates with the RNP complex and functions as

a signal to switch from the replication-transcription state to the viral assembly state [33]. These

mechanisms ensure that the machinery for viral transcription and replication is tightly

regulated.

Ebola Virus Glycoprotein GP1, 2

The fourth gene from the 3’ end of the filovirus genome encodes the viral envelope

GP1, 2 (a glycoprotein). Viral glycoprotein (GP1) is a heterotrimer that forms spikes on the

surface of the virus and is the only protein on the EBOV viral particle surface. Folding and

assembly of GP1 occurs independently of other viral proteins [34]. It is initially synthesized as

a precursor that is later cleaved by furin, a proprotein convertase, into GP1 (140kD) and GP2

(26kD) polypeptides that are linked by disulfide bonds [35]. The GP2 subunit contains two

heptad repeat regions that facilitate assembly of GP into trimers, a transmembrane anchor

sequence, and the fusion loop [34]. The heterotrimer (GP1 and GP2) assembles as a 450 KDa

complex on the surface of the nascent virion [36]. The GP1 subunit contains the cell surface

receptor binding site and a heavily glycosylated mucin domain. The residues responsible for

mediating interactions between GP1 and the host cell are within the 230-residue N-terminal

domain of GP1.

Page 17: Selection of functional RNA aptamers against Ebola

8

Deletion of the C-terminal mucin domain enhances viral infectivity in vitro [37-39].

This observation is explained by the crystal structure of EBOV GP (Figure2), in which the

receptor binding site of GP1 is masked by a glycan cap created by the mucin domain [36]. The

GP1 subunit responsible for interaction with the cell contains four different domains: base,

head, glycan cap and mucin like domain (MLD). The first three domains of GP1 are important

for expression and function of the prefusion state of the glycoprotein. The GP1 base forms a

hydrophobic, semicircular surface that interacts with the internal fusion loop and heptad repeat

of GP2, thereby acting as a clamp that holds the prefusion GP2 form. The head part of GP1 is

believed to contain the receptor binding domain (RBD) with a significant portion of this

domain exposed to the solvent.

The structure of the GP2 subunit, which is responsible for the fusion of viral and host

membrane, includes a hydrophobic internal fusion loop, two heptad repeats (HR1 and HR2)

and a CX6CC disulfide bond motif, membrane proximal external region and a transmembrane

anchor [40]. The internal fusion loop is wrapped around the outside of the GP trimer in the

pre-fusion state, with its hydrophobic side chains sequestered from the solvent due to packing

into a GP1 monomer.

Although Ebola virus species are antigenically distinct, the GP1 RBD is highly

conserved among them. The receptor binding domain of the Ebola viruses is masked by the

highly glycosylated domains of the mainly N-glycosylated glycan cap and a mucin like

domain, which is both N- and O-glycosylated [41]. Glycosyl groups protect proteins from

degradation but can also obstruct interaction of the glycosylated protein with another molecule.

Removal of the 7 glycan sites in the GP1 core had no significant effect on GP expression but

Page 18: Selection of functional RNA aptamers against Ebola

9

increased transduction compared with the wild type. Similar results were observed when the

mucin domain lacked its N-glycosylation [41]. These results demonstrated that the N linked

glycosyl groups on GP1 are not required for viral entry in Vero cells [41].

Ebola Virus Glycoprotein (GP1,2) Interaction With The Host Cell

Ebola viruses infect a broad range of cells and a host of proteins are known to enhance

viral entry into the host cells including the C-type lectins L-SIGN, DC-SIGN and the tyrosine

kinase receptor Axl [42-45]. Infectivity involves at least two recognition events 1) at the host

cell surface that mediates endocytosis and 2) in the endosome to mediate cytoplasmic entry.

The cell surface receptor for EBOV is believed to be the T-cell immunoglobulin and mucin

domain 1 (TIM-1) receptor [46]. The usual function of the TIM-1 receptor is to bind to

phosphatidyl serine (Ptd-Ser) on the surface of apoptotic cells and thereby facilitate

phagocytosis of these cells [47]. It has been shown that TIM-1 facilitates the entry of Dengue

Virus by directly interacting with the virion-associated phosphatidylserine [48].

Phosphatidylserine is the most highly represented anionic lipid in inner leaflet of plasma

membrane, where it usually is around 15-20% of anionic lipids [49]. Ebola Virus also gains

entry through interaction of the TIM-1 cell surface receptor and Ptd-Ser residue on the viral

envelope. The TIM-1 receptor enhances the internalization of pseudovirions that do not express

GP1, 2, confirming that TIM-1 and Ptd-Ser mediate the interaction for the viral entry into the

cell. [50]. After its initial interaction with the cell surface receptor, the virus enters the

endosomes through macropinocytosis. When the endosomes containing the virus particles

mature to late endosomes, the viral GP is further processed and it then interacts with the

Page 19: Selection of functional RNA aptamers against Ebola

10

endosomal receptor. However, although TIM-1 and DCSIGN have been identified as cellular

receptors for viral entry, the expression of TIM-1 and DCSIGN is not sufficient to make certain

cells permeable to Ebola virus entry. This result suggests that there may be distinct mechanisms

for Ebola mediated entry depending on which attachment factor is used for entry [51, 52].

Ebola virus requires endosomal processing before the final entry of the virus particle into the

cell. The protease, cathepsin B initiates entry which cleaves off the mucin domain and glycan

cap to expose the RBD [53-56].

The endosomal receptor for EBOV is believed to be the Niemann-Pick C1 (NPC1), a

lysosomal cholesterol transporter [57]. This conclusion is consistent with the observation that

Npc1−/− mice demonstrated significantly reduced viremia in comparison to wild type mice

when tested in a mouse model of Ebola disease [58]. NPC1 is a 13-pass transmembrane protein

found in the membranes of late endosomes and lysosomes of all cells [59]. It works together

with another protein, NPC2, to mediate the transport of cholesterol between cellular

compartments [60]. Binding site for C1 (NPC1) is exposed after the endosomal cysteine

proteases cleave EBOV GP1 to remove the heavily glycosylated C-terminal sequences, which

generates an entry intermediate comprised of the N terminal regions of GP1 and GP2 [54, 57].

The importance of the identified C1 (NPC1) binding site on GP1 was demonstrated with

virions that contained mutations in the C1 receptor binding residues, which are vital for viral

entry and for maintaining the conformation of the glycoprotein [57]. Recent studies have

shown that TIM-1 and NPC1-mediated interaction is responsible for viral fusion in the late

endosomes. Although, TIM-1 is primarily expressed on the cell surface, it is also found in early

endosomes and can cycle between the endosome and the cell surface [61]. A monoclonal

Page 20: Selection of functional RNA aptamers against Ebola

11

antibody (MabM224/1) against the TIM-1 receptor inhibits the internalization of Ebola virus

in Vero6 cells, which express the TIM-1 receptor. Thus, it is postulated that TIM-1 interacts

with the Ptd-Ser on Ebola virus, is taken into the endosome, and along with the cleaved

fragment of GP1, interacts with NPC-1 and initializes viral fusion with the late endosome lipid

membrane [62].

GP2-Mediated Viral and Host Membrane Fusion

The process of fusion of the viral and host cell membranes has a very high energy

barrier. To lower the activation barrier, viral proteins must undergo huge conformational

changes to generate sufficient released free energy. The first step for viral-host cell membrane

fusion in the endosome is to prime the envelope GP to a metastable conformation that can be

subsequently used for viral fusion. The crystal structure of EBOV GP1,2Δmucin (mucin

domain deleted) illustrates that the hydrophobic fusion residues are within the antiparallel β

scaffold and the structure is not conformationally restricted [63]. The N terminal of the fusion

loop has conformational freedom and is believed to aid fusion. In influenza and flaviviruses,

the main trigger for fusion is low pH, which results in protonation of histidine residues located

near positively charged residues in the prefusion state. These protonated histidines are found

in the post fusion conformation engaged in electrostatic interactions with the negatively

charged residues [64, 65]. The trigger for Ebola virus fusion is not well defined. Although the

low pH of the endosome is required for the cleavage of GP1,2 and for its subsequent interaction

with NPC1, the change in pH due to endosome maturation is not involved in triggering fusion.

However, it is unknown as yet if the cleavage of GP triggers the fusogenic conformation of the

Page 21: Selection of functional RNA aptamers against Ebola

12

virus [54] or the cleavage of GP by cathepsin B triggers another cellular factor that initiates

the fusion [55]. Even though the fusogenic trigger is unknown, it is evident from the crystal

structure that the GP2 hydrophobic patch, which clamps around GP1, must be released before

fusion. Comparing the pre- and post-fusion conformations of GP1,2 and GP2 respectively,

shows that the heptad repeat segments of GP2 unwinds from its prefusion state of a ring around

GP1, straightens and assembles into a 44 residue long helical rod like structure [63, 66, 67]. In

response to the rotational and translational movement of heptad repeat, the internal fusion loop

arranges itself to the top of the trimeric GP2 and adopts a 310 helical conformation, which is

stabilized by membrane interaction with hydrophobic residues in the fusion loop that mediate

penetration into the host membrane [68]. After formation of this intermediate conformation,

GP2 folds on itself bringing the two heptad repeats HR1 and HR2 close together to form a six-

helix bundle. This collapse of the two GP2 heptad repeats brings the host and viral membranes

closer causing the two bilayers to merge into a hemifusion stalk that eventually opens to a

fusion pore. (Figure 3).

Soluble Glycoprotein (sGP)

GP gene in the Ebola virus genome produces three different sized proteins: full length

676 aa GP1,2, pre-sGP which is 364 aa in length and a 298 residue small secreted protein

(ssGP) [69, 70]. These three proteins are produced because of transcriptional slippage of the

GP gene. GP1, 2 and ssGP are produced when slippage of the viral polymerase occurs at seven

consecutive uridines within the GP gene. With slippage of one U, an additional adenosine

residue is incorporated into the mRNA, which results in two sequential open reading frames

Page 22: Selection of functional RNA aptamers against Ebola

13

becoming a continuous reading frame that encodes full length GP1,2. sGP is produced from

the unedited version of the mRNA in which there is a stop codon at the end of the first reading

frame. Slippage along two Us and incorporation of two extra adenosines results in the

synthesis of ssGP, a secreted homodimer [69, 71]. The ratio of transcripts of these three

proteins have been shown to be (sGP:GP1.2:ssGP) of 71:24:5 in Vero E6 cells [71] suggesting

that sGP is the main viral protein product of the GP gene. Although the time of release of sGP

into the blood stream is unknown, sGP in 15 μL of serum, was readily detected by western blot

in comparison to GP1,2, which could not be detected [72].

Soluble glycoprotein (sGP) is encoded by the GP gene in all species of Ebola Virus. It

is initially synthesized as pre-GP which is shunted into the golgi complex via the signal

recognition complex. The sGP precursor undergoes post translational cleavage at the C-

terminus to yield the mature form of the protein [73]. After its proteolytic cleavage it forms a

103 KDa homodimer linked by two disulfide bonds at Cys 53 and Cys 306 [74, 75] with six

N-glycosylation sites [75]. sGP shares the first 295 amino acids with GP1, 2. A recently

determined cryo-EM structure of sGP showed that the monomeric structures of GP1,2 and

sGP are similar [Figure 4] [76]. Co-expression of sGP and GP2 by HEK293T cells resulted

in an sGP:GP2 protein complex that was recognized by the known neutralizing antibody KZ52

[77]. KZ52 is an antibody that was isolated from an Ebola survivor that binds to the interface

of GP1 and GP2 [63]. Using Circular Dichroism, Fluorescence Spectroscopy and MALLS it

was demonstrated that sGP, which is secreted as a disulfide linked homodimer, is mostly beta

sheet with 15% alpha helix content, and is highly stable at 37°C for up to 72 h [78].

Page 23: Selection of functional RNA aptamers against Ebola

14

Role Of sGP In Ebola Pathogenesis

sGP viral protein is secreted in large amounts by HEK293T cells and is present in the

blood of infected individuals [77], due to its high concentrations in blood various studies were

performed to determine its role in Ebola pathogenesis. One study suggested that sGP could

inactivate neutrophils by the binding to CD16b, which is a neutrophil specific Fc receptor III

[79]. However, these studies were challenged because the initial study overlooked the fact that

sGP could bind the neutrophil receptor indirectly through the Fc region of anti-sGP antibody

[80]. A later study confirmed that sGP immune complexes bound CD16b by demonstrating

that sGP complexed with an Fab fragment binds to neutrophils [81] These results argued

against the potential role of sGP as a neutrophil inactivator.

Another potential role of sGP was to promote apoptosis of B and T cells during viral

infection. Earlier reports had shown that lymphocyte apoptosis during Ebola infection could

be mediated through Fas/FasL interactions [82]. But, later it was demonstrated that sGP alone

or in combination with Fas ligands had no effect on frequency of Jurkat T cell death [82].

Investigators concluded that sGP does not induce apoptosis via extrinsic pathway. Thus, the

current evidence supports the conclusion that sGP is not involved in stimulating apoptosis of

lymphocytes.

A possible function of sGP in vascular dysregulation was also investigated, because

the expression of cellular adhesion molecules had earlier been shown to increase in the

presence of Ebola GP1,2. But, sGP and its degradation products failed to activate expression

of adhesion molecules in endothelial cells or to alter the integrity of endothelial cell barrier

[83]. Instead, when endothelial cells were treated with sGP and TNF-alpha (an inflammatory

Page 24: Selection of functional RNA aptamers against Ebola

15

signal that decreases endothelial barrier function), sGP restored barrier function, which

suggested an anti-inflammatory role for sGP [83]. Further studies explored the relationship

between the structure of sGP and its anti-inflammatory role. Mutations of Cys53Gly,

Cys306Gly, and the double mutant failed to restore the barrier function. Therefore, a fully

functional sGP homodimer is required for the protein to restore barrier function [75]. Overall,

these studies show that Ebola sGP might not have a role in the vascular deregulation but it

might have an anti-inflammatory function.

Immune modulation of host responses by sGP

Antibodies in the sera of human Ebola hemorrhagic fever survivors cross reacted with

sGP rather than GP1,2 [84]. Because sGP and GP1,2 are structurally similar they can be bound

by antibodies that recognize shared epitopes. However, due to the higher rate of production of

sGP in comparison to GP1,2, B cells that recognize sGP specific epitopes or shared epitopes

outcompete the B cells that recognize GP1,2 specific antibodies. This is believed to be the

reason for the biased population of antibodies that target sGP specific epitopes and GP1,2

shared epitope. Thus, the presence of high concentrations of sGP in the blood directs antibody

production towards sGP, which acts as a decoy to allow the viral infection to continue unabated

[85].

Vaccine and Therapeutics For Ebola

Until the 2014 Ebola outbreak no vaccines or therapeutics were available because the

infected population was very small and few companies were interested in the market. However,

Page 25: Selection of functional RNA aptamers against Ebola

16

with the magnitude of the outbreak in 2014, various vaccine formulations were proposed and

clinical trials were initiated. Clinical trials with rVSV-ZEBOV vaccine found that it was 70-

100% effective in providing protection therefore can be used in future outbreaks [86]. ChAd3-

ZEBOV, developed by GlaxoSmithkline in collaboration with the US national institute of

allergy and infectious diseases is still in clinical trial. The Novavax recombinant Ebola vaccine

is currently undergoing Phase II clinical trial. Other products in development include an oral

adenovirus (Vaxart), a different vesicular stomatitis virus candidate (Profectus Biosciences,

recombinant protein (Protein Sciences), DNA vaccine (Invovia) and a recombinant rabies

vaccine from (Jefferson University) [87].

Apart from the vaccines, there are many candidate therapeutics being evaluated

including small molecule inhibitors [88, 89] and siRNA based therapeutics [90]. ZMapp is a

cocktail of humanized antibodies that target three different sites on the Ebola glycoprotein

[91]. These three antibodies are produced by genetically modified tobacco plants and clinical

trials for Zmapp were concluded in 2016 [92]. The drug was well received by patients; however

it did not produce definitive results with the small number of patients in the trial. Therefore,

the FDA allowed Mapp Biopharmaceuticals to make Zmapp available to a larger number of

patients to obtain definitive answers.

Diagnostic Kits for EBOLA

WHO reported that, between 15th Feb 2015 and 19th April 2015, more than 26044

people were infected and around 10808 died from Ebola Virus [93]. As just discussed, several

potential drugs and vaccines are being tested, yet no FDA-approved treatment has yet been

Page 26: Selection of functional RNA aptamers against Ebola

17

commercialized. To minimize spread of the virus a means of early detection of Ebola virus is

needed so that infected individuals can be quarantined. By 2014 WHO had received 17

applications from diagnostic companies for devices to detect Ebola virus [94]. These included

reverse transcriptase polymerase chain reaction (RT-PCR) methods for detecting genomic

sequences, immunochromatographic strip formats and ELISAs. RT-PCR is the standard

diagnostic method for Ebola virus and is highly sensitive. However, it requires trained

personnel and specialized equipment in addition to high level biocontainment measures. In

addition, RT-PCR involves several steps during PCR amplification and requires constant

monitoring of the temperature. A constant power source is needed to limit the assay to specific

laboratories. The WHO approved a gold strip procedure for detecting Ebola virus. It is less

sensitive than RT-PCR and therefore would not detect an Ebola infection immediately after

the symptoms appear. It also has a 15% false positive rate, which limits its application as a

diagnostic kit. Current methods for Ebola detection dependent on the availability of necessary

reagents, require close contact with infected samples, and a constant supply of power. This

motivates interest in developing a robust, sensitive, and in-house diagnostic kit.

Aptamers

The term aptamer is derived from the latin word “aptus” which means “fit” and in

Greek the word meros, which means “particles”. Aptamers are short oligonucleotides with

sequences that enable them to bind to their selected targets with high affinity and specificity.

Aptamers have access to a repertoire of tertiary structural elements, including bulges, loops,

pseudoknots, and hairpin loops. Due to their varied structures and their capability to snugly

Page 27: Selection of functional RNA aptamers against Ebola

18

bind to the target molecule, the aptamer targets range from prokaryotic/eukaryotic cells, viral

particles, small molecules, and proteins. The interactions between the aptamer and target

consist mainly of a combination of stacking of aromatic rings, electrostatic and van der Waals

interactions and hydrogen bonding [95, 96].

SELEX (Systematic Evolution of Ligands through Exponential enrichment) is a

process that involves progressive siphoning of oligonucleotides (aptamers) through iterative

rounds of partitioning and amplification [97]. A randomized pool of RNA or ssDNA is

incubated with target molecules under specific buffer conditions. Oligonucleotides that bind

the target are separated from the unbound and are later amplified. The process is repeated until

the pool randomness is significantly reduced [96 168].

SELEX

Library Generation

The first step in SELEX process is to generate a randomized sequence oligonucleotide

library, which consists of ssDNA oligonucleotides comprised of a 20-60 nt random region

flanked by constant primer regions to be used for amplification. For selection of DNA aptamers

the synthesized pool can be directly used. Whereas for RNA aptamer selection, ssDNA pool

is converted to dsDNA and modified to include a T7RNA polymerase promoter using a primer.

The maximum number of random sequences in the library can be determined using 4n with n

being the number of positions in the random sequence. For our SELEX experiments we used

53 nucleotides in the random regions which translates to ~1032 possible sequences. However,

our starting pool size includes only 1015 oligonucleotides. It has been reported that the

Page 28: Selection of functional RNA aptamers against Ebola

19

probability of an aptamer for its target ranges from 10-13 to 10-14 with mean/mode probability

of 10-11 [98]. Therefore, a pool of 1015 complexity is frequently used for aptamer selection.

Things that need to be considered in the pool design are pool length, which will determine the

complexity of the pool, and nucleotide composition of the random sequences and constant

regions. Long oligonucleotides can be expected to adopt multiple secondary structures, which

might impede the correct folding of the desired binding motif. From studies to determine the

optimal length of the random region in oligonucleotide libraries it was suggested that longer

random pools (70-200nts) should be used to select for ribozymes, whereas shorter

oligonucleotides (<70nt) would serve the purpose for selecting aptamers [99]. Experiments

with pools containing different lengths of random region (16, 22, 26 50, 70 and 90) were used

to select for an aptamer that binds isoleucine, results showed that the optimum pool length for

isolating the desired aptamer was 50-70 nt [100]. A second important factor in designing pools

is the complexity of the pool. This depends on the ratio of the four nucleotides used to create

the ssDNA library, which can be determined in the manufacturing process [101, 102].

Chemically modified nucleotides can be used to introduce new features to aptamers

such as enhancing the sites of interaction with the target, improving the aptamer’s structural

stability or providing enhanced nuclease resistance [103, 104]. Two approaches can be used to

generate modified nucleotides. The first is to incorporate modified nucleotides at the time of

pool generation. However, this approach could be a problem if DNA or RNA polymerase will

not recognize the desired modified nucleotide as substrate. A second option is to modify the

already selected aptamer. However, a modification to an already selected aptamer might alter

the binding affinity or completely abolish binding [102].

Page 29: Selection of functional RNA aptamers against Ebola

20

Different techniques have been utilized to obtain modified aptamers conforming to a

desired function. For example, 2’F and 2’NH2 modifications in the 2’ ribose have been

included to improve the aptamer structural stability and nuclease resistance [105, 106].

Nucleotides containing photoactivable functional groups (e.g. 5 iodo,5-bromo and 4

thiouridine), which can covalently crosslink with a target protein, have also been used to

increase the affinity and specificity of the aptamers for their targets [107]. Modified aptamers

containing fluorescent groups have been used to study the binding of the aptamer to the protein

[108]. Amino acid side chain, like groups on 5-benzylaminocarbonyl-dU, 5-

napthylmethylaminocarbonyl-dU, 5-tryptaminocarbonyl-dU and 5-isobutylaminocarbonyl-

dU, confer a significant increase in chemical diversity, thereby increasing the chances of

interaction with proteins. For example, SOMAmers (Slow off rate modified aptamers) are

ssDNA aptamers with modified uracils and 5-methyl-dC [109].

Another important aspect of pool design is the design of constant regions. The sequence

of the constant region affects the amplification efficiency during PCR and RT (reverse

transcription) in case of RNA. The constant regions should be designed so as to not form stable

secondary structures [100]. Primers corresponding to the constant regions should also not form

strong secondary structures or include sequences that can cause primer dimers to form. To

improve amplification efficiency, the 3’ primer end should have WSS (W=T or A and S= G or

C) because polymerases can extend well with ACC at the 3’ end. A/T rich regions at the 3’

end can be used to avoid mispriming [110].

Page 30: Selection of functional RNA aptamers against Ebola

21

The selection process

SELEX process (Figure 5) starts with the incubation of the target with a pool of

oligonucleotides with random central sequences followed by a selection step in which the

bound oligonucleotide is separated from the unbound using a partitioning method. After this

separation for an RNA aptamer the selected RNA is reverse transcribed and further amplified

through PCR. In case of selection for a DNA aptamer the bound ssDNA is PCR-amplified.

Multiple rounds are performed in succession, which involve in vitro transcription for RNA

aptamer selection and separation of single stranded DNA for DNA aptamers. After multiple

rounds have been completed, the pools are analyzed by Next Gen Sequencing or Sanger

sequencing and evaluated for their ability to bind to the target with high affinity and specificity.

Aptamer Selection Methods

Filter capture

The filter capture method has been used extensively for selection of aptamers. The

principle factor in the method is the nonspecific binding of the target protein to the

nitrocellulose filter with the oligonucleotides passing through. Oligonucleotides bound to the

protein are trapped on the filter. This method has been widely successful but has the following

limitations 1) the efficiency of protein capture varies with the protein, 2) small molecules

cannot be captured on the membrane, and 3) certain aptamers bind to nitrocellulose

membranes in the absence of their protein target [111]. The filter capture method relies on the

true equilibrium binding of aptamers to the target protein in solution in comparison to other

methods in which either the target or the aptamer is immobilized on a solid support [112]. In

Page 31: Selection of functional RNA aptamers against Ebola

22

all these methods, including the filter capture method, the washing steps that follow the

partitioning can distort the equilibrium achieved in binding if the complex is characterized by

a high off-rate. However, the washing step is the shortest in the filter capture method compared

to others and thus least likely to influence the outcome. The filter capture method is also widely

used for evaluating the binding of the selected aptamers to the target protein [113].

Bead based SELEX

Immobilization of target molecules on a solid support is another method for selection

[114], various affinity tags (6X his, GST and MBP) can be used, in addition to coupling

chemistries such (amine, thiol or carboxyl). In bead based selection, target molecules can either

be pre-immobilized then incubated with the target library or the molecules can be incubated

with the target library and then captured by the affinity column. Various versions of affinity

chromatography have been used for selections such as 1) MEDUSA (Microplate based

enrichment device used for the selection of aptamers) [115], 2) the particle display method

[116] in which emulsion PCR is used to immobilize 105 copies of a single clonally amplified

ssDNA aptamer on each bead, the beads then incubated with fluorescent labeled target

molecules and sorted by flow cytometry, and 3) MonoLEX, which has been shown to generate

ssDNA aptamers in a single round that bind whole vaccinia virus particles [117].The

limitations of Bead-based SELEX are the restrictive interaction surface, requirement for

electronic instruments and flow pumps and increased non-specific interactions due to density-

dependent cooperative target binding [118].

Page 32: Selection of functional RNA aptamers against Ebola

23

Electrophoretic SELEX

EMSA (Electrophoretic Mobility shift assay) has long been used for downstream

applications during SELEX, mostly for evaluating the binding of potential aptamers to the

target protein. However, this method has also been used for aptamer selection [119]. Although

this method offers advantages such as equilibrium binding and separation of the bound and

unbound populations, it can be used with only some proteins and under limited buffer

conditions. As well, the volumes that can be accommodated in the gels used for separation are

too small to include the number of oligonucleotides required for the maximum possible

sequence diversity for SELEX. Therefore, it is wiser to use this method for downstream

applications.

A modification of EMSA is capillary electrophoresis which also relies on the

electrophoretic mobility of the aptamer, but separation is in the capillary as opposed to the gel

[120]. However, this method is limited to molecules that are within a certain size bracket and

that are charged so they can be resolved on the capillary.

Microfluidic SELEX

In the recent times, microfluidic SELEX has been used for selection of aptamers. There

are two different kinds of platforms 1) the target molecule is immobilized on micromagnetic

beads [121] and 2) the target molecule is encapsulated in sol-gel, gel like porous silica material

[122]. Microfluidic platforms allow fewer rounds of selection due to reduction in the available

target molecule along with function to perform continuous washing. This continuous washing

step allows removal of weak and nonspecific binders [123]. Since they are small, the use of

Page 33: Selection of functional RNA aptamers against Ebola

24

reagents is considerably reduced for microfluidic methods and the sol-gel based microfluidic

system can be utilized to select against multiple targets in one experiment. Even though the

advantages of this method are many, the usage of this method might be limited due to the

requirement for fabricated devices and electronic instruments. In sol gel based platform there

is also the chance of undesired effects of a chemical reaction during sol-gel formation on the

target molecule.

Cell SELEX

Cell-SELEX is used to select for aptamers against cells or cell fragments rather than

targeting a specific protein. In all SELEX procedures, negative selection against proteins or

cells other than the desired biomarker helps in gaining specificity towards the targeted

biomarker. Cell SELEX has been used to identify aptamers that can differentiate between

normal or diseased cell types [124, 125] and therefore the negative selective steps are generally

against normal cells. Cell SELEX has been used to select aptamers with therapeutic potential.

In terms of experiments, positive and negative selections can be done under conditions of

steady state equilibrium binding. Due to the possibility of nucleic acid aptamers being either

degraded or internalized during selection, Cell-SELEX protocols usually employ shorter

incubation periods than necessary for reaching equilibrium. Since cell surfaces also display

non-target cellular proteins, it is difficult to achieve very high specificity towards the target

molecules. TECS SELEX addresses this problem by overexpressing the desired target protein

for selections. An aptamer that recognizes the TGF-β type III receptor with nanomolar affinity

was successfully selected using this method [126]. Another complication of Cell-SELEX is

Page 34: Selection of functional RNA aptamers against Ebola

25

that it is prone to artifacts resulting from the presence of dead cells in the selection population

Cell-SELEX protocols use FACS to separate bound and unbound aptamers [127]. This enables

sorting of live and dead cells along with selecting cells with bound aptamers [128].

High Throughput Sequencing and Data Analysis

Incorporation of high throughput sequencing into the SELEX protocol for the analysis

of enriched aptamer libraries and identification of candidate aptamers has been the most

informative recent change in SELEX applications. Previously, enriched aptamer pools were

cloned into a plasmid and a few hundred individual clones were sequenced to identify high

affinity aptamers. In contrast, high throughput sequencing analysis has a readout of over a

hundred million sequences. Candidate aptamers can be selected at much earlier rounds by

studying the enrichment of an oligonucleotide with a particular sequence across various

rounds, whereas conventional cloning-based aptamer identification required multiple rounds

until an ensured enriched population was observed [129]. Illumina platforms are a preferred

platform for NGS as they offer a significantly higher number of sequence reads with read

lengths, which would cover the random regions and also the constant regions of most libraries

[130, 131]. A large number of reads helps in gaining a more comprehensive insight into the

sequence and structural features of the aptamers selected and a broad perspective on the

progress of each SELEX process. Information gathered through NGS has also helped to

predict secondary structure of the RNA aptamer and to identify appropriate truncations that

yielded a better binding aptamer [132]. Given the ability of NGS to sequence multiple SELEX

rounds simultaneously, it can be used to monitor the selection process as it proceeds. For

Page 35: Selection of functional RNA aptamers against Ebola

26

example, in a recent study SELEX pools derived from a genomic library selected against the

E.coli Hfq protein were run in parallel with a control SELEX (Neutral SELEX) that did not

include the target during the selection steps and was subjected to the same PCR and RT

treatments. The analysis of various rounds revealed that selection with the target progressed

independently of the Neutral SELEX, with the former resulting in substantial enrichment of

certain sequences whereas with the latter there was no enrichment of sequences [133].

It is widely accepted now that even highly enriched libraries contain many thousands

of potential aptamer sequences, many of them present in the populations in lower numbers. It

has also become more evident that the most abundant oligonucleotides obtained in the final

SELEX pools might not have the sequences for the highest target affinities. Other than

selection for affinity to the target, reasons for high abundance of an oligonucleotide with a

sequence in a selected pool could be PCR amplification bias or affinity of the oligonucleotide

with that sequence to the partitioning matrix. Therefore, the fold enrichment over one or more

SELEX rounds rather than abundance is a much better determinant of high affinity aptamers

[131]. Tracking fold-enrichment is a good way to identify aptamers especially if the starting

library is biased.

The current capability of generating an immense amount of data and its contribution to

better understanding the SELEX process has resulted in high throughput sequencing being

used by most aptamer research groups to analyze the SELEX results. Despite its popularity,

the development of specific bioinformatics tools for identification of candidate aptamers from

the huge amounts of sequencing data is limited. Current bioinformatics tools based on

clustering identify top clusters and the aptamer population can be traced to identify the

Page 36: Selection of functional RNA aptamers against Ebola

27

enriched sequences. Current tools also allow sorting of the putative aptamer sequences via their

sequence structure motifs [134-136]. However, bioinformatics tools are still needed that better

predict the target binding aptamer candidates based on structure or sequence classifications.

Aptamers vs. Antibodies and The Current State Of The Market

Antibodies, have been widely used for various purposes. Apart from being used in the

lab for research, they are used as therapeutics drugs, are also functionalized on biosensors for

detection purposes. Although, antibodies have been fundamental to several applications, there

are limitations associated with its usability. Synthesis of an antibody is time consuming and

there is always an issue of reproducibility between batches. Neutralizing antibodies have been

traditionally used to prevent the spread of viral infections. But, no neutralizing antibodies have

yet been commercialized for EBOV. Also, antibodies used in passive immunity must be

“humanized” to prevent their immunogenicity. Thus, even if there were an effective anti-

EBOV antibody available, it would be very expensive and time consuming to humanize the

antibody. Antibodies due to their short shelf life, require proper storage conditions, therefore

when used for a detection platform, conditions amenable for preserving antibodies is needed.

In addition, to perform techniques like RT-PCR or ELISAs trained personnel is required.

Usage of antibodies adds up to additional costs and require infrastructure, regions affected by

Ebola viruses have poor infrastructure therefore storage of these drugs and kits require

additional costs and efforts.

Aptamers have been established as a potent alternative for antibodies. Aptamers have

been selected against several targets to function either as a therapeutic drug or as a biosensor.

Page 37: Selection of functional RNA aptamers against Ebola

28

Aptamers have notable advantages over antibodies that include non-immunogenicity, in-vitro

production, and stability. Aptamers can be rapidly reconstituted and the in vivo

pharmacokinetics and tissue distribution can be tuned by chemical modifications including

conjugation to high molecular weight molecules such as PEG. Conjugation can increase the

half-life of an aptamer from 40 min to 50 h. Modifications such as with fluorine in the 2’

position also increase stability in serum [137].

Even though aptamers were discovered in 1990 they still haven’t effectively transcended from

research labs into the market full-fledged even though they hold many advantages over

antibodies (Table 1). Only one aptamer-based product Macugen has cleared clinical trials and

is available in the market. The huge investment of pharmaceutical companies in the antibody

market may be partly responsible for the resistance against aptamers. Even though few

aptamers are in the market, the research on aptamers hasn’t receded. A handful of companies

have been developing and marketing aptamers as antibody-like binding reagents for research

and development. These companies are Soma Logic (Boulder, CO, USA), OTC Biotech (San

Antonio, TX, USA), Aptagen (Jacobus, PA, USA), Base Pair Biotechnologies (Houston, TX,

USA), NeoVentures Biotechnologies (London, ON, Canada), Aptamer Sciences (Pohang,

Korea). A few companies marketed aptamers as components of assay kits or concentrating

devices, such as NeoVentures Ochratoxin A. Soma Logic , a company at the forefront of

aptamer research, has been pushing its technology, SOMAmers and SOMAscan aptamer

platform sensors, in the diagnostic market [138].

Page 38: Selection of functional RNA aptamers against Ebola

29

Aptamers In The Diagnosis Of Viral Infections

Proper diagnosis of any viral infection is important for preventing its spread. Symptoms

of an acute infection are at first non-diagnostic and associated with a variety of viral infections.

To prevent its spread, it is imperative that the infectious agent be identified in the early stages.

Aptamer development to date has been concentrated mostly on the well-known viral diseases

such as HIV, Influenza virus, SARS and HPV [155-158]. Little research has been done on a

wider variety of other viral diseases, including Ebola virus. Very few aptamers have been

reported that bind Ebola viral proteins or are directed towards blocking an antiviral function

[159, 160].

There is a wide range of aptasensor options designed based on different principles for

detection of viral particles. An aptasensor was designed to detect avian influenza virus, based

on the principle of quartz crystal microbalance [161]. Gold microelectrodes with impedimetric

Table 1: Advantages of aptamers

Advantages References

Bind to a wide range of targets including

metal ions, metabolites, proteins, viruses &

mammalian cells

[139-144]

Aptamers can be synthesized chemically,

reducing batch to batch variation and can be

used as tools for drug loading and antidote

application.

[145]

Aptamers generate little to no

immunogenicity in therapeutic applications.

[140, 146-148]

Aptamers can be conjugated to molecules

such as drugs, carriers, toxins and siRNA for

therapeutic applications.

[148-151]

Aptamers can be conjugated with imaging

probes for molecular imaging applications

[152-154]

Page 39: Selection of functional RNA aptamers against Ebola

30

properties allows one to distinguish between active and inactive forms of virus. Using specific

DNA aptamers viable virus was detected by using the impedimetric sensor [162]. Other

approaches to detect viral antigens include an RNA aptamer sensor to the HCV core antigen

in which fluorescent dye ( Cy3) conjugated RNA aptamer was the means of detection.[163]

In another study an aptamer to viral glycoprotein E2 was selected using CS SELEX and

demonstrated to inhibit HCV infection in Huh7.5.1 cells [144].

Comparison of aptamer based tests with other analytical methods

Most current methods to detect virus infections rely on Enzyme linked immunosorbent

assay (ELISA) or RT-PCR [161, 164, 165]. However, ELISA is reported to have low

sensitivity and very high rates of false positives [166-168]. It also cannot be broadly applied

because of the need for specific antibodies that are difficult to obtain for several viral diseases

[169]. RT-PCR can be performed without application of antibodies and can detect the viral

genome efficiently. It is also a sensitive method. Although the advantages are many for RT-

PCR, it requires expensive enzymes and trained personnel to do these analyses.

Aptamers could be used to circumvent bottlenecks in the current detection systems for

viral diseases. Properly developed and authenticated aptamers could efficiently differentiate

between infected and non-infected cells.

On a Hydrogel coated QCM (Quartz crystal microbalance aptasensor) biosensor, anti-

H5 antibodies were coated in parallel with anti-H5 aptamers. Although, same concentration of

antibody was coated as the aptamer, the detection limit of H5N1 with anti-H5 antibody was

0.128U whereas with the aptamer the detection limit was 0.0128U. [161, 167]. In addition to

Page 40: Selection of functional RNA aptamers against Ebola

31

providing higher sensitivity, detection using aptamers takes less time than with antibodies or

PCR. For example, to detect avian influenza virus the time for detection by ELISA or qPCR

would be 3 hours and 5 hours respectively, whereas with an aptasensor it only takes 1.5 hours

[165-168].

Aptamers In Preventing The Fusion Of Virus Particle To The Host Cell

Apart from diagnosis, another obvious function of an aptamer could be as a therapeutic.

An aptamer could act as a therapeutic agent by preventing the interaction between the viral

proteins and the host cell receptor proteins. Many aptamers have been selected that inhibit viral

entry. An RNA aptamer (B40) that binds HIV gp120 inhibits viral interaction with T cell co-

receptor CCR5. Application of the aptamer resulted in a decreased concentration of p24

antigen in the supernatants from virus-infected cultures of human peripheral mononuclear

blood cells as measured by ELISA. Further characterization of the aptamer showed that it binds

to the core conserved region of gp120 [170-172].

E2 protein is a potential target for creating blocking aptamers against HCV infection.

E2 is a coreceptor of human CD81, which is presented on hepatocytes and B lymphocytes.

ZE2, a DNA aptamer selected against, E2 blocked E2 binding of a wide range of HCV

serotypes to Huh7.5.1, human established cells line of hepatocellular carcinoma cells. The

presence of the aptamer reduced viral RNA levels was tested by qPCR. [144].

A DNA aptamer A22 to the influenza virus HA protein, blocked influenza virus entry

into MDCK (Madin-Darby Canine Kidney). Cell viability increased in proportion to A22

aptamer concentrations from 50 to 100 pM. The therapeutic potential of this aptamer was also

Page 41: Selection of functional RNA aptamers against Ebola

32

demonstrated in an animal model in which mice infected with A/Texas/1/7 influenza strain and

simultaneously treated with the aptamer A22, lost weight slower that the non-treated control

group [173]. C7-35M, a DNA aptamer that binds to receptor binding region of hemagglutinin

binds to the virus particles and inhibits the binding of the virus to the host cell receptors. [174].

Similar studies tested the effect of aptamers against human cytomegalovirus infections.

Two RNA aptamers that recognize HCMV viral glycoproteins B and H, reduced the infectivity

of cytomegalovirus in a human foreskin fibroblast cell line culture [175]. Aptamers against

HSV-1 bound viral glycoprotein D, a ligand of Nectin-1, which mediates HSV entry into the

host cell. Reduced infectious potential of HSV-1 was observed in proportion to the dose of this

aptamer. The treatments showed no evidence of cytotoxicity and the aptamer could distinguish

between HSV-1 and HSV-2 strains [176].

In my thesis, data from SELEX experiments to select 2’Fluoro pyrimidine modified

RNA aptamers against Ebola glycoproteins GP1,2 and sGP is presented.

Page 42: Selection of functional RNA aptamers against Ebola

34

Figure 1: Ebola Virus Particle.

Filamentous Ebola virus particle, Nucleocapsid comprises of RNA polymerase (L),

Nucleoprotein (NP), VP35, VP30. Major and minor matrix proteins laying beneath the viral

membrane VP40 and VP24. Anchored glycoprotein (GP) on the surface of the virus particles.

Reprinted by permission from [177].

Page 43: Selection of functional RNA aptamers against Ebola

35

Figure 2: X ray crystallography determined structure of GP1,2 ΔmucinΔtm.

The structure of Ebola virus glycoprotein bound to neutralizing antibody KZ52. GP1,2 is a

heterotrimer expressed on the surface of the viral membrane. The cartoon depicts the mucin

domain that shrouds the protein along with glycan cap. The red region is supposed to be the

receptor binding domain. Reprinted by permission from [63].

Page 44: Selection of functional RNA aptamers against Ebola

36

Figure 3: Prefusion and and postfusion conformational switch of GP1,2 glycoprotein.

The cartoon depicts the interaction of GP1,2 with host cell receptors followed by

internalization via micropinocytosis. Once the virus is internalized into the endosomes, low

pH dependent proteases cleave the glycan cap and mucin domain to expose the receptor

binding domain 19-KDa core bonded to GP2. The receptor binding domain interacts with a

receptor in the endosomes and this interaction allows a conformational change within GP2.

This conformational change allows GP2 to release its internal fusion loop (IFL) and insert into

host cell membrane. Interaction between the heptad repeats 2 and 1 brings host cell and viral

membrane into contact and forms a hemifusion stalk. The hemifusion stalk causes the

formation of a fusion pore. Reprinted by permission from [36].

Page 45: Selection of functional RNA aptamers against Ebola

37

Figure 4: CryoEM determined structures of full length GP1,2 glycoprotein and soluble

glycoprotein (sGP). Cryo EM determined quaternary structure of GP1,2. The figure also

depicts the residues of N linked glycosylation and mucin domain sites.Cryo EM determined

structure of sGP shows a parallel homodimer connected by disulfide bonds at the N and C

terminus. The monomer structures of GP1,2 and sGP are similar. Reprinted by permission

from [76]

Page 46: Selection of functional RNA aptamers against Ebola

38

Figure 5: SELEX protocol for the selection of RNA aptamer that binds the target protein.

A DNA pool with a complexity of 1015 was PCR amplified, followed by transcription to

prepare an RNA pool. The RNA pool was incubated with target protein and non-binders were

separated by capture on a nitrocellulose filter. Retained RNA was eluted and reverse

transcribed followed by PCR to amplify the selected sequences. This process was performed

for 8-10 times before submitting for NGS.

Page 47: Selection of functional RNA aptamers against Ebola

39

References

1. Kuhn, J.H., et al., Proposal for a revised taxonomy of the family Filoviridae:

classification, names of taxa and viruses, and virus abbreviations. Archives of virology, 2010.

155(12): p. 2083-2103.

2. Elliott, L.H., M.P. Kiley, and J.B. McCormick, Descriptive analysis of Ebola

virus proteins. Virology, 1985. 147(1): p. 169-176.

3. Noda, T., et al., Ebola virus VP40 drives the formation of virus-like filamentous

particles along with GP. Journal of virology, 2002. 76(10): p. 4855-4865.

4. Takada, A., Filovirus tropism: cellular molecules for viral entry. Frontiers in

microbiology, 2012. 3.

5. Dolnik, O., L. Kolesnikova, and S. Becker, Filoviruses: Interactions with the

host cell. Cell Mol Life Sci, 2008. 65(5): p. 756-76.

6. Benton, A. and K.Y. Dionne, International Political Economy and the 2014

West African Ebola Outbreak. African Studies Review, 2015. 58(01): p. 223-236.

7. Gunaratne, N.D., Ebola Virus and the Threat of Bioterrorism, The. Fletcher F.

World Aff., 2015. 39: p. 63.

8. Geisbert, T.W., et al., Pathogenesis of Ebola hemorrhagic fever in cynomolgus

macaques: evidence that dendritic cells are early and sustained targets of infection. The

American journal of pathology, 2003. 163(6): p. 2347-2370.

9. BREMEN, J., et al., The epidemiology of Ebola hemorrhagic fever in Zaire.

1976, SR Pattyn. New York, Elsevier/North-Holland Biomedical Press.

Page 48: Selection of functional RNA aptamers against Ebola

40

10. Geisbert, T.W., et al., Pathogenesis of Ebola hemorrhagic fever in primate

models: evidence that hemorrhage is not a direct effect of virus-induced cytolysis of

endothelial cells. The American journal of pathology, 2003. 163(6): p. 2371-2382.

11. Amara, A. and J. Mercer, Viral apoptotic mimicry. Nature Reviews

Microbiology, 2015. 13(8): p. 461-469.

12. Biermann, M., et al., Surface code—biophysical signals for apoptotic cell

clearance. Physical biology, 2013. 10(6): p. 065007.

13. Morizono, K. and I.S. Chen, Role of phosphatidylserine receptors in enveloped

virus infection. Journal of virology, 2014. 88(8): p. 4275-4290.

14. Prins, K.C., W.B. Cárdenas, and C.F. Basler, Ebola virus protein VP35 impairs

the function of interferon regulatory factor-activating kinases IKKε and TBK-1. Journal of

virology, 2009. 83(7): p. 3069-3077.

15. Basler, C.F., et al., The Ebola virus VP35 protein functions as a type I IFN

antagonist. Proceedings of the National Academy of Sciences, 2000. 97(22): p. 12289-12294.

16. Basler, C.F., et al., The Ebola virus VP35 protein inhibits activation of

interferon regulatory factor 3. Journal of virology, 2003. 77(14): p. 7945-7956.

17. Cárdenas, W.B., et al., Ebola virus VP35 protein binds double-stranded RNA

and inhibits alpha/beta interferon production induced by RIG-I signaling. Journal of virology,

2006. 80(11): p. 5168-5178.

18. Leung, D.W., et al., Structure of the Ebola VP35 interferon inhibitory domain.

Proceedings of the National Academy of Sciences, 2009. 106(2): p. 411-416.

Page 49: Selection of functional RNA aptamers against Ebola

41

19. Leung, D.W., et al., Structural basis for dsRNA recognition and interferon

antagonism by Ebola VP35. Nature structural & molecular biology, 2010. 17(2): p. 165-172.

20. Ivashkiv, L.B. and L.T. Donlin, Regulation of type I interferon responses.

Nature Reviews Immunology, 2014. 14(1): p. 36-49.

21. Leung, L.W., et al., Ebola virus VP24 binds karyopherin α1 and blocks STAT1

nuclear accumulation. Journal of virology, 2006. 80(11): p. 5156-5167.

22. Shabman, R.S., et al., The Ebola virus VP24 protein prevents hnRNP C1/C2

binding to karyopherin α1 and partially alters its nuclear import. Journal of Infectious Diseases,

2011. 204(suppl 3): p. S904-S910.

23. Kolesnikova, L., et al., Ultrastructural organization of recombinant Marburg

virus nucleoprotein: comparison with Marburg virus inclusions. Journal of virology, 2000.

74(8): p. 3899-3904.

24. Watanabe, S., T. Noda, and Y. Kawaoka, Functional mapping of the

nucleoprotein of Ebola virus. Journal of virology, 2006. 80(8): p. 3743-3751.

25. Noda, T., et al., Assembly and budding of Ebolavirus. PLoS Pathog, 2006. 2(9):

p. e99.

26. Johnson, R.F., et al., Ebola virus VP35-VP40 interaction is sufficient for

packaging 3E-5E minigenome RNA into virus-like particles. Journal of virology, 2006. 80(11):

p. 5135-5144.

27. Becker, S., et al., Interactions of Marburg virus nucleocapsid proteins.

Virology, 1998. 249(2): p. 406-417.

Page 50: Selection of functional RNA aptamers against Ebola

42

28. Mühlberger, E., et al., Three of the four nucleocapsid proteins of Marburg virus,

NP, VP35, and L, are sufficient to mediate replication and transcription of Marburg virus-

specific monocistronic minigenomes. Journal of virology, 1998. 72(11): p. 8756-8764.

29. Mühlberger, E., et al., Comparison of the transcription and replication strategies

of Marburg virus and Ebola virus by using artificial replication systems. Journal of virology,

1999. 73(3): p. 2333-2342.

30. Whelan, S., J. Barr, and G. Wertz, Transcription and replication of

nonsegmented negative-strand RNA viruses, in Biology of Negative Strand RNA Viruses: The

Power of Reverse Genetics. 2004, Springer. p. 61-119.

31. Shabman, R.S., et al., An upstream open reading frame modulates Ebola virus

polymerase translation and virus replication. PLoS Pathog, 2013. 9(1): p. e1003147.

32. Biedenkopf, N., et al., Phosphorylation of Ebola virus VP30 influences the

composition of the viral nucleocapsid complex impact on viral transcription and replication.

Journal of Biological Chemistry, 2013. 288(16): p. 11165-11174.

33. Watanabe, S., et al., Ebola virus (EBOV) VP24 inhibits transcription and

replication of the EBOV genome. Journal of Infectious Diseases, 2007. 196(Supplement 2): p.

S284-S290.

34. Sanchez, A., et al., Biochemical analysis of the secreted and virion

glycoproteins of Ebola virus. J Virol, 1998. 72(8): p. 6442-7.

35. Volchkov, V.E., et al., Processing of the Ebola virus glycoprotein by the

proprotein convertase furin. Proc Natl Acad Sci U S A, 1998. 95(10): p. 5762-7.

Page 51: Selection of functional RNA aptamers against Ebola

43

36. Lee, J.E. and E.O. Saphire, Ebolavirus glycoprotein structure and mechanism

of entry. Future Virol, 2009. 4(6): p. 621-635.

37. Manicassamy, B., et al., Comprehensive analysis of ebola virus GP1 in viral

entry. J Virol, 2005. 79(8): p. 4793-805.

38. Jeffers, S.A., D.A. Sanders, and A. Sanchez, Covalent modifications of the

ebola virus glycoprotein. J Virol, 2002. 76(24): p. 12463-72.

39. Brindley, M.A., et al., Ebola virus glycoprotein 1: identification of residues

important for binding and postbinding events. J Virol, 2007. 81(14): p. 7702-9.

40. Gallaher, W.R., Similar structural models of the transmembrane proteins of

Ebola and avian sarcoma viruses. Cell, 1996. 85(4): p. 477-478.

41. Lennemann, N.J., et al., Comprehensive functional analysis of N-linked glycans

on Ebola virus GP1. MBio, 2014. 5(1): p. e00862-13.

42. Ji, X., et al., Mannose-binding lectin binds to Ebola and Marburg envelope

glycoproteins, resulting in blocking of virus interaction with DC-SIGN and complement-

mediated virus neutralization. J Gen Virol, 2005. 86(Pt 9): p. 2535-42.

43. Takada, A., et al., Human macrophage C-type lectin specific for galactose and

N-acetylgalactosamine promotes filovirus entry. J Virol, 2004. 78(6): p. 2943-7.

44. Simmons, G., et al., DC-SIGN and DC-SIGNR bind ebola glycoproteins and

enhance infection of macrophages and endothelial cells. Virology, 2003. 305(1): p. 115-23.

45. Brindley, M.A., et al., Tyrosine kinase receptor Axl enhances entry of Zaire

ebolavirus without direct interactions with the viral glycoprotein. Virology, 2011. 415(2): p.

83-94.

Page 52: Selection of functional RNA aptamers against Ebola

44

46. Kondratowicz, A.S., et al., T-cell immunoglobulin and mucin domain 1 (TIM-

1) is a receptor for Zaire Ebolavirus and Lake Victoria Marburgvirus. Proc Natl Acad Sci U S

A, 2011. 108(20): p. 8426-31.

47. Kobayashi, N., et al., TIM-1 and TIM-4 glycoproteins bind phosphatidylserine

and mediate uptake of apoptotic cells. Immunity, 2007. 27(6): p. 927-40.

48. Meertens, L., et al., The TIM and TAM families of phosphatidylserine receptors

mediate dengue virus entry. Cell Host Microbe, 2012. 12(4): p. 544-57.

49. Adu-Gyamfi, E., et al., Host Cell Plasma Membrane Phosphatidylserine

Regulates the Assembly and Budding of the Ebola Virus. Journal of virology, 2015: p. JVI.

01087-15.

50. Moller-Tank, S., et al., Role of the phosphatidylserine receptor TIM-1 in

enveloped-virus entry. Journal of virology, 2013. 87(15): p. 8327-8341.

51. Alvarez, C.P., et al., C-type lectins DC-SIGN and L-SIGN mediate cellular

entry by Ebola virus in cis and in trans. Journal of virology, 2002. 76(13): p. 6841-6844.

52. Lasala, F., et al., Mannosyl glycodendritic structure inhibits DC-SIGN-

mediated Ebola virus infection in cis and in trans. Antimicrobial agents and chemotherapy,

2003. 47(12): p. 3970-3972.

53. Dube, D., et al., The primed ebolavirus glycoprotein (19-kilodalton GP1, 2):

sequence and residues critical for host cell binding. Journal of virology, 2009. 83(7): p. 2883-

2891.

54. Chandran, K., et al., Endosomal proteolysis of the Ebola virus glycoprotein is

necessary for infection. Science, 2005. 308(5728): p. 1643-5.

Page 53: Selection of functional RNA aptamers against Ebola

45

55. Schornberg, K., et al., Role of endosomal cathepsins in entry mediated by the

Ebola virus glycoprotein. Journal of virology, 2006. 80(8): p. 4174-4178.

56. Hood, C.L., et al., Biochemical and structural characterization of cathepsin L-

processed Ebola virus glycoprotein: implications for viral entry and immunogenicity. Journal

of virology, 2010. 84(6): p. 2972-2982.

57. Miller, E.H., et al., Ebola virus entry requires the host-programmed recognition

of an intracellular receptor. EMBO J, 2012. 31(8): p. 1947-60.

58. Herbert, A.S., et al., Niemann-Pick C1 Is Essential for Ebolavirus Replication

and Pathogenesis In Vivo. mBio, 2015. 6(3): p. e00565-15.

59. Davies, J.P. and Y.A. Ioannou, Topological analysis of Niemann-Pick C1

protein reveals that the membrane orientation of the putative sterol-sensing domain is identical

to those of 3-hydroxy-3-methylglutaryl-CoA reductase and sterol regulatory element binding

protein cleavage-activating protein. Journal of Biological Chemistry, 2000. 275(32): p. 24367-

24374.

60. Infante, R.E., et al., NPC2 facilitates bidirectional transfer of cholesterol

between NPC1 and lipid bilayers, a step in cholesterol egress from lysosomes. Proceedings of

the National Academy of Sciences, 2008. 105(40): p. 15287-15292.

61. Balasubramanian, S., et al., TIM family proteins promote the lysosomal

degradation of the nuclear receptor NUR77. Science signaling, 2012. 5(254): p. ra90.

62. Kuroda, M., et al., Interaction between TIM-1 and NPC1 Is Important for

Cellular Entry of Ebola Virus. Journal of virology, 2015. 89(12): p. 6481-6493.

Page 54: Selection of functional RNA aptamers against Ebola

46

63. Lee, J.E., et al., Structure of the Ebola virus glycoprotein bound to an antibody

from a human survivor. Nature, 2008. 454(7201): p. 177-182.

64. Bullough, P.A., et al., Structure of influenza haemagglutinin at the pH of

membrane fusion. Nature, 1994. 371(6492): p. 37-43.

65. Kampmann, T., et al., The role of histidine residues in low-pH-mediated viral

membrane fusion. Structure, 2006. 14(10): p. 1481-1487.

66. Weissenhorn, W., et al., Crystal structure of the Ebola virus membrane fusion

subunit, GP2, from the envelope glycoprotein ectodomain. Molecular cell, 1998. 2(5): p. 605-

616.

67. Malashkevich, V.N., et al., Core structure of the envelope glycoprotein GP2

from Ebola virus at 1.9-Å resolution. Proceedings of the National Academy of Sciences, 1999.

96(6): p. 2662-2667.

68. Freitas, M.S., et al., Structure of the Ebola fusion peptide in a membrane-

mimetic environment and the interaction with lipid rafts. Journal of Biological Chemistry,

2007. 282(37): p. 27306-27314.

69. Sanchez, A., et al., The virion glycoproteins of Ebola viruses are encoded in

two reading frames and are expressed through transcriptional editing. Proceedings of the

National Academy of Sciences, 1996. 93(8): p. 3602-3607.

70. VOLCHKOV, V.E., et al., GP mRNA of Ebola Virus Is Edited by the Ebola

Virus Polymerase and by T7 and Vaccinia Virus Polymerases 1. Virology, 1995. 214(2): p.

421-430.

Page 55: Selection of functional RNA aptamers against Ebola

47

71. Mehedi, M., et al., A new Ebola virus nonstructural glycoprotein expressed

through RNA editing. Journal of virology, 2011. 85(11): p. 5406-5414.

72. Sanchez, A., et al., Detection and molecular characterization of Ebola viruses

causing disease in human and nonhuman primates. Journal of Infectious Diseases, 1999.

179(Supplement 1): p. S164-S169.

73. Volchkov, V.E., et al., Processing of the Ebola virus glycoprotein by the

proprotein convertase furin. Proceedings of the National Academy of Sciences, 1998. 95(10):

p. 5762-5767.

74. Barrientos, L.G., et al., Disulfide bond assignment of the Ebola virus secreted

glycoprotein SGP. Biochemical and biophysical research communications, 2004. 323(2): p.

696-702.

75. Falzarano, D., et al., Structure‐Function Analysis of the Soluble Glycoprotein,

sGP, of Ebola Virus. Chembiochem, 2006. 7(10): p. 1605-1611.

76. Pallesen, J., et al., Structures of Ebola virus GP and sGP in complex with

therapeutic antibodies. Nature microbiology, 2016. 1: p. 16128.

77. Iwasa, A., M. Shimojima, and Y. Kawaoka, sGP serves as a structural protein

in Ebola virus infection. Journal of Infectious Diseases, 2011. 204(suppl 3): p. S897-S903.

78. Barrientos, L.G., et al., Secreted Glycoprotein from Live Zaire ebolavirus—

Infected Cultures: Preparation, Structural and Biophysical Characterization, and

Thermodynamic Stability. Journal of Infectious Diseases, 2007. 196(Supplement 2): p. S220-

S231.

Page 56: Selection of functional RNA aptamers against Ebola

48

79. Yang, Z.-y., et al., Distinct cellular interactions of secreted and transmembrane

Ebola virus glycoproteins. Science, 1998. 279(5353): p. 1034-1037.

80. Sui, J. and W.A. Marasco, Evidence against Ebola virus sGP binding to human

neutrophils by a specific receptor. Virology, 2002. 303(1): p. 9-14.

81. Maruyama, T., et al., Ebola virus, neutrophils, and antibody specificity.

Science, 1998. 282(5390): p. 843-843.

82. Wauquier, N., et al., Human fatal zaire ebola virus infection is associated with

an aberrant innate immunity and with massive lymphocyte apoptosis. PLoS Negl Trop Dis,

2010. 4(10): p. e837.

83. Wahl-Jensen, V.M., et al., Effects of Ebola virus glycoproteins on endothelial

cell activation and barrier function. Journal of virology, 2005. 79(16): p. 10442-10450.

84. Maruyama, T., et al., Recombinant human monoclonal antibodies to Ebola

virus. Journal of Infectious Diseases, 1999. 179(Supplement 1): p. S235-S239.

85. Mohan, G.S., et al., Antigenic subversion: a novel mechanism of host immune

evasion by Ebola virus. PLoS pathogens, 2012. 8(12): p. e1003065.

86. Henao-Restrepo, A.M., et al., Efficacy and effectiveness of an rVSV-vectored

vaccine in preventing Ebola virus disease: final results from the Guinea ring vaccination, open-

label, cluster-randomised trial (Ebola Ça Suffit!). The Lancet, 2017. 389(10068): p. 505-518.

87. World Health Organization. Ebola vaccines, therapies, and diagnostics. 2015;

Available from: http://www.who.int/medicines/emp_ebola_q_as/en/.

88. Côté, M., et al., Small molecule inhibitors reveal Niemann-Pick C1 is essential

for Ebola virus infection. Nature, 2011. 477(7364): p. 344-348.

Page 57: Selection of functional RNA aptamers against Ebola

49

89. Warren, T.K., et al., Protection against filovirus diseases by a novel broad-

spectrum nucleoside analogue BCX4430. Nature, 2014. 508(7496): p. 402-405.

90. Geisbert, T.W., et al., Postexposure protection of non-human primates against

a lethal Ebola virus challenge with RNA interference: a proof-of-concept study. The Lancet,

2010. 375(9729): p. 1896-1905.

91. Murin, C.D., et al., Structures of protective antibodies reveal sites of

vulnerability on Ebola virus. Proceedings of the National Academy of Sciences, 2014. 111(48):

p. 17182-17187.

92. PREVAIL, I. and M.-N.P.I.S. Team, A Randomized, Controlled Trial of ZMapp

for Ebola Virus Infection. The New England journal of medicine, 2016. 375(15): p. 1448.

93. World Health Organization. Ebola Situation Report. 2015; Available from:

http://apps.who.int/ebola/current-situation/ebola-situation-report-22-april-2015-0.

94. Butler, D., Ebola experts seek to expand testing. Nature, 2014. 516(7530): p.

154-155.

95. Ellington, A.D. and J.W. Szostak, In vitro selection of RNA molecules that bind

specific ligands. nature, 1990. 346(6287): p. 818-822.

96. Hermann, T. and D.J. Patel, Adaptive recognition by nucleic acid aptamers.

Science, 2000. 287(5454): p. 820-825.

97. Gopinath, S.C.B., Methods developed for SELEX. Analytical and bioanalytical

chemistry, 2007. 387(1): p. 171-182.

98. Gold, L., et al., Diversity of oligonucleotide functions. Annual review of

biochemistry, 1995. 64(1): p. 763-797.

Page 58: Selection of functional RNA aptamers against Ebola

50

99. Sabeti, P.C., P.J. Unrau, and D.P. Bartel, Accessing rare activities from random

RNA sequences: the importance of the length of molecules in the starting pool. Chemistry &

biology, 1997. 4(10): p. 767-774.

100. Legiewicz, M., et al., Size, constant sequences, and optimal selection. Rna,

2005. 11(11): p. 1701-1709.

101. Famulok, M. and G. Mayer, Aptamers as tools in molecular biology and

immunology, in Combinatorial Chemistry in Biology. 1999, Springer. p. 123-136.

102. Kulbachinskiy, A., Methods for selection of aptamers to protein targets.

Biochemistry (Moscow), 2007. 72(13): p. 1505-1518.

103. Eaton, B.E., L. Gold, and D.A. Zichi, Let's get specific: the relationship

between specificity and affinity. Chemistry & biology, 1995. 2(10): p. 633-638.

104. Kusser, W., Chemically modified nucleic acid aptamers for in vitro selections:

evolving evolution. Reviews in Molecular Biotechnology, 2000. 74(1): p. 27-38.

105. Jayasena, S.D., Aptamers: an emerging class of molecules that rival antibodies

in diagnostics. Clinical chemistry, 1999. 45(9): p. 1628-1650.

106. Nimjee, S.M., C.P. Rusconi, and B.A. Sullenger, Aptamers: an emerging class

of therapeutics. Annu. Rev. Med., 2005. 56: p. 555-583.

107. Jensen, K.B., et al., Using in vitro selection to direct the covalent attachment of

human immunodeficiency virus type 1 Rev protein to high-affinity RNA ligands. Proceedings

of the National Academy of Sciences, 1995. 92(26): p. 12220-12224.

108. Nutiu, R. and Y. Li, Aptamers with fluorescence-signaling properties. Methods,

2005. 37(1): p. 16-25.

Page 59: Selection of functional RNA aptamers against Ebola

51

109. Gold, L., et al., Aptamer-based multiplexed proteomic technology for

biomarker discovery. PloS one, 2010. 5(12): p. e15004.

110. Pollard, J., S.D. Bell, and A.D. Ellington, Design, synthesis, and amplification

of DNA pools for construction of combinatorial pools and libraries. Current Protocols in

Molecular Biology, 2000: p. 24.2. 1-24.2. 24.

111. Shi, H., et al., Evolutionary dynamics and population control during in vitro

selection and amplification with multiple targets. RNA, 2002. 8(11): p. 1461-1470.

112. Ozer, A., J.M. Pagano, and J.T. Lis, New technologies provide quantum

changes in the scale, speed, and success of SELEX methods and aptamer characterization.

Molecular Therapy—Nucleic Acids, 2014. 3(8): p. e183.

113. Hall, K.B. and J.K. Kranz, Nitrocellulose filter binding for determination of

dissociation constants, in RNA-protein interaction protocols. 1999, Springer. p. 105-114.

114. Nieuwlandt, D., M. Wecker, and L. Gold, In vitro selection of RNA ligands to

substance P. Biochemistry, 1995. 34(16): p. 5651-5659.

115. Latulippe, D.R., et al., Multiplexed microcolumn-based process for efficient

selection of RNA aptamers. Analytical chemistry, 2013. 85(6): p. 3417-3424.

116. Wang, J., et al., Particle Display: A Quantitative Screening Method for

Generating High‐Affinity Aptamers. Angewandte Chemie International Edition, 2014. 53(19):

p. 4796-4801.

117. Nitsche, A., et al., One-step selection of Vaccinia virus-binding DNA aptamers

by MonoLEX. BMC biotechnology, 2007. 7(1): p. 48.

Page 60: Selection of functional RNA aptamers against Ebola

52

118. Ozer, A., et al., Density-dependent cooperative non-specific binding in solid-

phase SELEX affinity selection. Nucleic acids research, 2013. 41(14): p. 7167-7175.

119. Tsai, R.Y. and R.R. Reed, Identification of DNA recognition sequences and

protein interaction domains of the multiple-Zn-finger protein Roaz. Molecular and cellular

biology, 1998. 18(11): p. 6447-6456.

120. Mendonsa, S.D. and M.T. Bowser, In vitro evolution of functional DNA using

capillary electrophoresis. Journal of the American Chemical Society, 2004. 126(1): p. 20-21.

121. Huang, C.-J., et al., Integrated microfluidic system for rapid screening of CRP

aptamers utilizing systematic evolution of ligands by exponential enrichment (SELEX).

Biosensors and Bioelectronics, 2010. 25(7): p. 1761-1766.

122. Park, S.-m., et al., Selection and elution of aptamers using nanoporous sol-gel

arrays with integrated microheaters. Lab on a Chip, 2009. 9(9): p. 1206-1212.

123. Oh, S.S., et al., Improving aptamer selection efficiency through volume

dilution, magnetic concentration, and continuous washing in microfluidic channels. Analytical

chemistry, 2011. 83(17): p. 6883-6889.

124. Morris, K.N., et al., High affinity ligands from in vitro selection: complex

targets. Proceedings of the National Academy of Sciences, 1998. 95(6): p. 2902-2907.

125. Shangguan, D., et al., Aptamers evolved from live cells as effective molecular

probes for cancer study. Proceedings of the National Academy of Sciences, 2006. 103(32): p.

11838-11843.

Page 61: Selection of functional RNA aptamers against Ebola

53

126. Ohuchi, S.P., T. Ohtsu, and Y. Nakamura, Selection of RNA aptamers against

recombinant transforming growth factor-β type III receptor displayed on cell surface.

Biochimie, 2006. 88(7): p. 897-904.

127. Raddatz, M.S.L., et al., Enrichment of Cell‐Targeting and Population‐Specific

Aptamers by Fluorescence‐Activated Cell Sorting. Angewandte Chemie, 2008. 120(28): p.

5268-5271.

128. Avci-Adali, M., et al., Pitfalls of cell-systematic evolution of ligands by

exponential enrichment (SELEX): existing dead cells during in vitro selection anticipate the

enrichment of specific aptamers. Oligonucleotides, 2010. 20(6): p. 317-323.

129. Cheng, C., et al., In vivo SELEX for Identification of Brain-penetrating

Aptamers. Molecular Therapy—Nucleic Acids, 2013. 2(1): p. e67.

130. Hoon, S., et al., Aptamer selection by high-throughput sequencing and

informatic analysis. Biotechniques, 2011. 51(6): p. 413-416.

131. Cho, M., et al., Quantitative selection of DNA aptamers through microfluidic

selection and high-throughput sequencing. Proceedings of the National Academy of Sciences,

2010. 107(35): p. 15373-15378.

132. Berezhnoy, A., et al., Isolation and optimization of murine IL-10 receptor

blocking oligonucleotide aptamers using high-throughput sequencing. Molecular Therapy,

2012.

133. Zimmermann, B., et al., Monitoring genomic sequences during SELEX using

high-throughput sequencing: neutral SELEX. PLoS One, 2010. 5(2): p. e9169.

Page 62: Selection of functional RNA aptamers against Ebola

54

134. Beier, R., E. Boschke, and D. Labudde, New strategies for evaluation and

analysis of SELEX experiments. BioMed research international, 2014. 2014.

135. Ditzler, M.A., et al., High-throughput sequence analysis reveals structural

diversity and improved potency among RNA inhibitors of HIV reverse transcriptase. Nucleic

acids research, 2012: p. gks1190.

136. Hoinka, J., et al., Identification of sequence–structure RNA binding motifs for

SELEX-derived aptamers. Bioinformatics, 2012. 28(12): p. i215-i223.

137. Ruckman, J., et al., 2'-Fluoropyrimidine RNA-based aptamers to the 165-amino

acid form of vascular endothelial growth factor (VEGF165). Inhibition of receptor binding and

VEGF-induced vascular permeability through interactions requiring the exon 7-encoded

domain. J Biol Chem, 1998. 273(32): p. 20556-67.

138. Bruno, J.G., Predicting the Uncertain Future of Aptamer-Based Diagnostics and

Therapeutics. Molecules, 2015. 20(4): p. 6866-6887.

139. Kawakami, J., et al., In vitro selection of aptamers that act with Zn 2+. Journal

of Inorganic Biochemistry, 2000. 82(1): p. 197-206.

140. Neff, C.P., et al., An aptamer-siRNA chimera suppresses HIV-1 viral loads and

protects from helper CD4(+) T cell decline in humanized mice. Sci Transl Med, 2011. 3(66):

p. 66ra6.

141. Bruno, J.G., et al., Competitive FRET-aptamer-based detection of

methylphosphonic acid, a common nerve agent metabolite. Journal of fluorescence, 2008.

18(5): p. 867-876.

Page 63: Selection of functional RNA aptamers against Ebola

55

142. Ruckman, J., et al., 2′-Fluoropyrimidine RNA-based aptamers to the 165-amino

acid form of vascular endothelial growth factor (VEGF165) Inhibition of receptor binding and

VEGF-induced vascular permeability through interactions requiring the exon 7-encoded

domain. Journal of Biological Chemistry, 1998. 273(32): p. 20556-20567.

143. Tang, Z., et al., Generating aptamers for recognition of virus-infected cells.

Clinical chemistry, 2009. 55(4): p. 813-822.

144. Chen, F., et al., CS-SELEX generates high-affinity ssDNA aptamers as

molecular probes for hepatitis C virus envelope glycoprotein E2. PloS one, 2009. 4(12): p.

e8142.

145. M Bompiani, K., et al., Antidote control of aptamer therapeutics: the road to a

safer class of drug agents. Current pharmaceutical biotechnology, 2012. 13(10): p. 1924-1934.

146. Group, E.S., Preclinical and phase 1A clinical evaluation of an anti-VEGF

pegylated aptamer (EYE001) for the treatment of exudative age-related macular degeneration.

Retina, 2002. 22(2): p. 143-152.

147. Foy, J.W.-D., et al., Local tolerance and systemic safety of pegaptanib sodium

in the dog and rabbit. Journal of Ocular Pharmacology and Therapeutics, 2007. 23(5): p. 452-

466.

148. Meng, L., et al., Targeted delivery of chemotherapy agents using a liver cancer-

specific aptamer. PLoS One, 2012. 7(4): p. e33434.

149. Liu, Z., et al., Novel HER2 aptamer selectively delivers cytotoxic drug to

HER2-positive breast cancer cells in vitro. J Transl Med, 2012. 10(1): p. 148.

Page 64: Selection of functional RNA aptamers against Ebola

56

150. Subramanian, N., et al., Target-specific delivery of doxorubicin to

retinoblastoma using epithelial cell adhesion molecule aptamer. Molecular vision, 2012. 18: p.

2783.

151. Zhang, M.-Z., et al., Targeted quantum dots fluorescence probes functionalized

with aptamer and peptide for transferrin receptor on tumor cells. Nanotechnology, 2012.

23(48): p. 485104.

152. Hong, H., et al., Molecular imaging with nucleic acid aptamers. Current

medicinal chemistry, 2011. 18(27): p. 4195.

153. Song, Y., et al., Selection of DNA aptamers against epithelial cell adhesion

molecule for cancer cell imaging and circulating tumor cell capture. Analytical chemistry,

2013. 85(8): p. 4141-4149.

154. Talbot, L.J., et al., Pharmacokinetic characterization of an RNA aptamer against

osteopontin and demonstration of in vivo efficacy in reversing growth of human breast cancer

cells. Surgery, 2011. 150(2): p. 224-230.

155. Cho, S.-J., et al., Novel system for detecting SARS coronavirus nucleocapsid

protein using an ssDNA aptamer. Journal of bioscience and bioengineering, 2011. 112(6): p.

535-540.

156. Toscano-Garibay, J.D., M.L. Benítez-Hess, and L.M. Alvarez-Salas, Isolation

and characterization of an RNA aptamer for the HPV-16 E7 oncoprotein. Archives of medical

research, 2011. 42(2): p. 88-96.

157. Tombelli, S., et al., Aptamer-based biosensors for the detection of HIV-1 Tat

protein. Bioelectrochemistry, 2005. 67(2): p. 135-141.

Page 65: Selection of functional RNA aptamers against Ebola

57

158. Negri, P., et al., Direct optical detection of viral nucleoprotein binding to an

anti-influenza aptamer. Analytical chemistry, 2012. 84(13): p. 5501-5508.

159. Binning, J.M., et al., Development of RNA aptamers targeting Ebola virus

VP35. Biochemistry, 2013. 52(47): p. 8406-8419.

160. Huang, Z., X. Wang, and G. Gao, Analyses of SELEX-derived ZAP-binding

RNA aptamers suggest that the binding specificity is determined by both structure and

sequence of the RNA. Protein & cell, 2010. 1(8): p. 752-759.

161. Wang, R. and Y. Li, Hydrogel based QCM aptasensor for detection of avian

influenzavirus. Biosensors and Bioelectronics, 2013. 42: p. 148-155.

162. Labib, M., et al., Aptamer-based viability impedimetric sensor for viruses.

Analytical chemistry, 2012. 84(4): p. 1813-1816.

163. Lee, S., et al., Chip-based detection of hepatitis C virus using RNA aptamers

that specifically bind to HCV core antigen. Biochemical and biophysical research

communications, 2007. 358(1): p. 47-52.

164. Fletcher, S.J., et al., Toward specific detection of Dengue virus serotypes using

a novel modular biosensor. Biosensors and Bioelectronics, 2010. 26(4): p. 1696-1700.

165. Bai, H., et al., A SPR aptasensor for detection of avian influenza virus H5N1.

Sensors, 2012. 12(9): p. 12506-12518.

166. Luo, Q., et al., An indirect sandwich ELISA for the detection of avian influenza

H5 subtype viruses using anti-hemagglutinin protein monoclonal antibody. Veterinary

microbiology, 2009. 137(1): p. 24-30.

Page 66: Selection of functional RNA aptamers against Ebola

58

167. Dhumpa, R., et al., Rapid detection of avian influenza virus in chicken fecal

samples by immunomagnetic capture reverse transcriptase–polymerase chain reaction assay.

Diagnostic microbiology and infectious disease, 2011. 69(3): p. 258-265.

168. Sidoti, F., et al., Development of real time RT-PCR assays for detection of type

A influenza virus and for subtyping of avian H5 and H7 hemagglutinin subtypes. Molecular

biotechnology, 2010. 44(1): p. 41-50.

169. Szpechciński, A. and A. Grzanka, [Aptamers in clinical diagnostics]. Postepy

biochemii, 2005. 52(3): p. 260-270.

170. Cohen, C., et al., An aptamer that neutralizes R5 strains of HIV-1 binds to core

residues of gp120 in the CCR5 binding site. Virology, 2008. 381(1): p. 46-54.

171. Dey, A.K., et al., An aptamer that neutralizes R5 strains of human

immunodeficiency virus type 1 blocks gp120-CCR5 interaction. Journal of virology, 2005.

79(21): p. 13806-13810.

172. DEY, A.K., et al., Structural characterization of an anti-gp120 RNA aptamer

that neutralizes R5 strains of HIV-1. Rna, 2005. 11(6): p. 873-884.

173. Jeon, S.H., et al., A DNA aptamer prevents influenza infection by blocking the

receptor binding region of the viral hemagglutinin. Journal of Biological Chemistry, 2004.

279(46): p. 48410-48419.

174. Choi, S.K., et al., DNA aptamers against the receptor binding region of

hemagglutinin prevent avian influenza viral infection. Molecules and cells, 2011. 32(6): p.

527-533.

Page 67: Selection of functional RNA aptamers against Ebola

59

175. WANG, J., H. JIANG, and F. LIU, In vitro selection of novel RNA ligands that

bind human cytomegalovirus and block viral infection. Rna, 2000. 6(04): p. 571-583.

176. Gopinath, S.C., K. Hayashi, and P.K. Kumar, Aptamer that binds to the gD

protein of herpes simplex virus 1 and efficiently inhibits viral entry. Journal of virology, 2012.

86(12): p. 6732-6744.

177. Moller-Tank, S. and W. Maury, Ebola virus entry: a curious and complex series

of events. PLoS pathogens, 2015. 11(4): p. e1004731.

178. Ksiazek, T.G., et al., ELISA for the detection of antibodies to Ebola viruses.

The Journal of infectious diseases, 1999. 179(Supplement_1): p. S192-S198.

179. Su, S., et al., Diagnostic strategies for Ebola virus detection. The Lancet

Infectious Diseases, 2016. 16(3): p. 294-295.

180. Kanai, A., K. Tanabe, and M. Kohara, Poly (U) binding activity of hepatitis C

virus NS3 protein, a putative RNA helicase. FEBS letters, 1995. 376(3): p. 221-224.

Page 68: Selection of functional RNA aptamers against Ebola

60

CHAPTER 2

A 2’FY-RNA EPITOPE FORMS AN APTAMER FOR EBOLAVIRUS sGP:

SELECTION AND CHARACTERIZATION

Shambhavi Shubhama,b, Jan Hoinkac, Emma Swansona, Teresa M. Przytyckac,

Wendy Mauryd, Soma Banerjee a,b and Marit Nilsen-Hamiltona,b

aIowa State University, Ames IA, bAptalogic Inc., Ames IA, cNational Institutes of Health,

Bethesda, DC, dUniversity of Iowa, Iowa City IA

Short title: A 2’FY-RNA Epitope Selectively Binds Ebola Virus sGP

Experimental contributions to this manuscript:

SELEX against sGP, identification of sequence structure motif, binding assays to determine

poly U and 2’FY modification dependent high affinity binding, truncation of parent aptamer

to determine the RNA binding epitope, antisense hybridization assay and binding assays to

determine the Kd of aptamer 39SGP1A against sGP were performed by Shambhavi Shubham

Analysis of Next Gen Sequencing data and the identification of top 10 clusters was done by

Jan Hoinka.

Emma Swanson repeated the binding assays of 39SGP1A against sGP and performed binding

assays to determine the affinity of 39SGP1A against serum proteins.

Soma Banerjee determined the binding affinity of 39SGP1A against sGP using an Aptamer

Sandwich Binding Assay that she developed.

Page 69: Selection of functional RNA aptamers against Ebola

61

ABSTRACT

Nucleic acid aptamers have advantages over antibodies for incorporation into sensing

devices. These advantages include stability to storage in the dehydrated state, structural

stability for repetitive use, and compatibility with many sensing/transducing mechanisms.

Such capability is important for detecting Ebola virus (EBOV) for which epidemics frequently

start in locations that lack sophisticated equipment. sGP (soluble glycoprotein), a nonstructural

Ebola viral glycoprotein, is an EBOV biomarker that is secreted in abundance in the blood

stream even during the early stages of infection. Here, we report the selection of 39SGP1A, a

2’flouro pyrimidine (2’FY)-modified RNA aptamer that specifically binds to sGP and does not

bind human serum albumin, the most abundant protein in human blood. We demonstrate by

computational and biochemical analysis that the recognition motif of 39SGP1A is a novel

polypyrimidine-rich sequence. The replacement of the F group in the 2’ position of the ribose

with an OH group resulted in complete loss of affinity for sGP. Computational structural

modeling suggests that this motif could form a compact surface for protein interaction for

which modification by fluorine increases molecular hydrophobicity.

Page 70: Selection of functional RNA aptamers against Ebola

62

INTRODUCTION

Ebola virus, in the family of Filoviridae, is a highly virulent enveloped RNA virus that

causes hemorrhagic fever with 50-90% fatality and for which early detection is the key to

controlling epidemics such as have occurred in the recent past. The Ebola virus GP gene

encodes two glycoproteins with the surface exposed trimeric glycoprotein (GP1,2) being the

minor product. The major translated product is sGP (soluble glycoprotein), which is conserved

amongst the five Ebola virus species (1,2). sGP shares its first 295 amino acids with GP1,2

but has a unique 28 amino acid carboxy terminus. Monomers are disulphide linked (at C53 and

C306) in a parallel orientation and appear in the blood as homodimers. It is N glycosylated at

6 positions and C-mannosylated at W288. Due to their sequence homology sGP and GP1,2

have similar monomeric structures (3). Although, the role of sGP in pathogenesis is not yet

established, antibodies elicited during infection cross react with sGP and GP1,2. This

observation supports role for sGP as a decoy for antibodies against GP1,2 on the virus particle

(4). Consistent with this hypothesis is the observation that sGP is secreted in abundance in the

early stages of infection. Thus, it is an excellent choice as an Ebola virus biomarker.

Despite its clear advantage as a biomarker for Ebola virus infections, sGP has been so

far overlooked for this purpose. Here we describe the application of SELEX (5-7) combined

with Next Gen Sequencing (NGS) and analysis by APTAGUI to select a 2’FY-RNA aptamer

(39SGP1A) with high affinity and specificity for sGP. Aptamer 39SGP1A binds sGP tightly

with a Kd of 27nM. By contrast, 39SGP1A does not bind human serum albumin. The

requirement for structure in the binding of 39SGP1A to sGP is evident by the lack of affinity

for sGP of the RNA (2’OH) version of this aptamer, which distinguishes the 39SGP1A motif

Page 71: Selection of functional RNA aptamers against Ebola

63

from the RNA recognition domain of the polypyrimidine tract binding protein (8-10). 2’F

modification, which provides resistance against endonucleases (11) also results in

conformational differences from the 2’OH RNA. The 2’F-modified ribose sugar tends to adopt

the C3’ endo conformation and stabilizes the RNA structure (12). In addition, the inclusion of

fluorine modifications increases molecular hydrophobicity. Here, we describe the aptamer and

characterize its interaction with sGP, identifying a novel 2’FY-RNA binding domain as a poly-

2’F-U loop with GAGC in the stem.

RESULTS

Selection of 2’FY-RNAs with High Affinity for sGP

2’FY-RNAs with high affinity for sGP were selected by SELEX from an

oligonucleotide pool, consisting of a 53nt random region sequence flanked by 25 bases 5’ and

3’ constant regions with a starting complexity of 1015 molecules of RNA. The protocol

followed a mathematically defined approach of starting with a relatively high molar ratio of

2:1 (RNA: sGP) followed by harmonic reductions in the sGP concentration (13) to increase

the stringency of selections (Fig.1B). To eliminate non-specific sequences binding to the

nitrocellulose membrane, negative selections were performed at rounds 2 and 4. After 9 rounds

of selection, enrichment of the pools was analyzed by nitrocellulose filter capture assay and

EMSA. The binding curve of the initial pool did not reach saturation, whereas the binding

curve of the final pool was biphasic with saturation achieved at ~100 nM and ~1uM

respectively suggesting selection of high and low affinity RNA binders (data not shown). The

filter binding results were corroborated by the results from electrophoretic mobility shift assay

Page 72: Selection of functional RNA aptamers against Ebola

64

(EMSA) results using the pools from the 1st and 9th rounds RNA pools. The 1st, 4th, 6th and 9th

rounds pools were given for Next Gen Sequencing Analysis and simultaneously 9th round pool

was sub-cloned and sequenced by TOPO-TA cloning.

Sequencing Analysis and Identification of Conserved Sequence Motifs

APTA GUI (14), an open source graphical user interface was used for the Next Gen

Sequencing data analysis. APTAGUI provides essential information regarding the quality of

the data by separating the reads into several rounds and extracting randomized sequences.

APTA GUI also generates enrichment ratios across the selection cycles, which helps in

selecting prospective sequences for further analysis. In addition to individual sequences it also

provides information about various aptamer clusters and sequence families. The APTAGUI

platform allows the user to study the mutation count on each individual sequence. Due to its

extensive analysis of the aptamer families, it greatly facilitates the identification of aptamers.

Next Gen Sequencing Analysis (Figs. 1A, C, D). The top 10 Sequence clusters obtained by

TOPO TA cloning were identical to the Top 10 clusters obtained from Next Gen Sequencing

analysis.

Analysis of the sequences in the top 10 clusters from NGS with MEME (15) identified

a conserved sequence motif of a stretch of U’s and GAGC (Fig. 1E). No other sequence

structure motifs were found in other enriched clusters. In all the secondary structures predicted

for the sequence motif by MFold (16) the poly 2’F-U’s were in a loop followed by GAGC in

a stem.

Page 73: Selection of functional RNA aptamers against Ebola

65

The poly2’F-U-GAGC Sequence Motif Identifies A High Affinity Aptamer

The affinities of 2’FY-RNA oligonucleotides 5183, 5177, 5179 and 5182 containing

the poly 2’F-U-GAGC sequence motif were compared with oligonucleotides 4789, 5181 from

the top sequence cluster that did not contain the motif. In the initial titrations, we observed

saturation at 100nM for all sequences containing U rich motif suggesting high affinity binding.

The results, demonstrated for 2’FY-RNA-5183 and 2’FY-RNA-4789 revealed high affinities

(Kd = 20-50 nM) for oligonucleotides containing the poly 2’F-U-GAGC sequence motif and

low affinities (Kd = 0.5 to 1uM) for sequences that did not contain the motif (Fig. 2 A-D). That

the 2’FY modification is essential for 2’FY-RNA-5183 binding to sGP is demonstrated by the

observation that the same sequence in the form of an RNA (with 2’OH) did not bind sGP (Fig.

2A). sGP binding by 2’FY-RNA 5183 was confirmed by electrophoretic mobility shift assay

(EMSA, Fig. 2E).

The specificity of 2’FY-RNA-5183 binding to sGP in the presence and absence of an

unrelated RNA aptamer (RNA-569) that recognizes mouse Lcn2 (17). RNA-569, in 10 molar

excess, did not compete for 2’FY-RNA-5183 binding to sGP (Fig. 2E). These results are

consistent with the hypothesis that 2’FY-RNA-5183 binds sGP with specificity and high

affinity. Cluster containing 2’FY-RNA-5183 was examined for base frequency at each position

by APTAGUI to establish a “mutation rate” through the selections (14). 5’ end stem loop,

which contains the poly 2’F-U-GAGC sequence motif, was identified as the most stable

sequence. Truncations of the 2’FY-RNA-5183 were tested for binding sGP (Fig. 3A, B, D).

These results showed that the high affinity binding element in 2’FY-RNA-5183 contains the

poly 2’F-U-GAGC sequence motif.

Page 74: Selection of functional RNA aptamers against Ebola

66

A 2’FY-RNA-5199, a 39nt truncated version of 5199 containing the poly 2’F-U-GAGC

sequence motif that binds sGP with Kd of 27nM was given the aptamer name of 39SGP1A.

Further truncation to 2’FY-RNA-5199 containing only 6 U’s within the loop (Oligo 5015) had

lower affinity for sGP, supporting the hypothesis that the structure of the U rich loop, rather

than just the presence of a series of 2’FY modified Us, is important for binding (Fig. 3F). We

also tested the binding of 39SGP1A with sGP by performing UV crosslinking and observed a

clear mobility shift when the RNA and protein were combined. We confirmed that the shifted

band was an RNA-protein photo-adduct by digestion with Proteinase K (Fig. 3C). We also

tested involvement of the poly 2’F-U loop by hybridizing it with antisense oligonucleotide

DNA-5196. The presence of a 10-fold molar excess of DNA-5196 over 39SGP1A resulted in

50% reduction in the sGP-RNA shifted band after crosslinking. These results demonstrate that

39SGP1A is a 2’FY-RNA aptamer that bind sGP with high affinity.

How Many Epitopes On sGP Are Recognized By The Selected 2’FY-RNA

Oligonucleotides?

Several RNA oligonucleotide families were identified from the NGS results that are

not homologous in sequence. Whereas members of the family identified by the poly 2’F-U-

GAGC sequence motif bound sGP with high affinity, members of other families bound with

lower affinity (example in Fig. 2). To establish if these oligonucleotides bound to the same or

different protein epitopes, we tested the ability of a representative of another family to compete

for binding of 39sGP1A with sGP.

Page 75: Selection of functional RNA aptamers against Ebola

67

The 39SGP1A aptamer, 2’FY RNA oligonucleotides 4789, 5179 and oligonucleotide

6011, a DNA oligo that binds sGP, were used as competitors of 32P-39SGP1A for binding sGP.

Competition was observed with all the oligonucleotides, which varied in affinity for sGP. This

observation suggests that sGP has one favored nucleic acid binding epitope and the high

affinity and low affinity binding is determined by the tertiary structure of the RNA (Fig.4).

Specificity of 39SGP1A

To establish if 39SGP1A is suitable as a sensor for early diagnosis of sGP in patient

serum, we tested its affinity for human serum albumin (HSA). HSA is highly abundant in

blood, reaching concentrations up to 300 μM. The 2’FY-RNA oligonucleotide-5183, from

which 39SGP1A was derived, bound HSA with a Kd of 400 nM (Fig. 5A). However, once

isolated from the remainder of 2’FY-RNA-5183, 39SGP1A did not bind HSA up to 68 μM

(Fig. 5B). 39SGP1A was also found not to bind two other highly abundant serum proteins,

alpha 2 macroglobulin and fibrinogen (Fig. 5C).

DISCUSSION

In this study, we identified an RNA aptamer that binds sGP with high affinity and

specificity. sGP is a non-structural Ebola virus 70 KDa glycoprotein that forms a parallel

homodimer linked by disulfide bonds, which is responsible for immune modulation during

infection (4). The choice of sGP as a SELEX target was motivated by the fact that it is secreted

in abundance in blood early during infection, which makes it an excellent biomarker for Ebola

infection (2). An aptamer that binds sGP could be incorporated onto a biosensor platform for

Page 76: Selection of functional RNA aptamers against Ebola

68

early infection detection. 39SGP1A is the first RNA aptamer reported to be specific for sGP.

For our selections, we began with a complexity of 1015 2’FY modified RNA oligonucleotides.

2’F ribose modification provides nuclease resistance and has widely been used in aptamer

selections. After eight rounds of selection we compared the binding of the final pool with the

starting pool by the filter capture assay. The initial RNA pool showed no evidence of saturation

at 5 μM, whereas in the final round RNA pool the sGP titration indicated the presence of two

groups of oligonucleotides identified by high and low affinities. We used conventional TOPO

TA cloning for high throughput sequencing of the 8th round and performed Next Gen

Sequencing on 1st, 4th, 6th and 8th Round pools. Clusters obtained from high throughput

sequencing utilizing TOPO TA cloning were identical to the clusters identified by Next Gen

Sequencing.

To identify potential aptamers, we searched for conserved motifs within the top 10

clusters and identified a consensus sequence of poly U and GAGC. In the predicted secondary

structures, the poly U stretch of 18 nts resided in the loop region followed by GAGC’s in the

stem region. Oligonucleotides containing the poly U motif demonstrated high affinity binding

(Kds = 20-34 nM), whereas oligonucleotides from the top ten selected clusters that did not

contain the poly U motif bound sGP with low affinity. Certain cellular proteins to bind to

poly(U) rich RNA sequences (18-20). However, the selected oligonucleotides in this study

contained 2’F- modified pyrimidines, which alters the structure by holding the sugar ribose in

the C3’ endo conformation. This conformation is usually favored in RNA duplexes, increasing

the helical stability. When oligonucleotide 5183 was prepared with 2’OH uracil and 2’F

Page 77: Selection of functional RNA aptamers against Ebola

69

cytosine or 2’ OH cytosine binding to sGP was lost, showing that 2’F-U modification is

essential for this oligonucleotide to bind sGP with high affinity.

A comparison of selected aptamers from libraries of RNA molecules containing

2'amino and 2'fluoro deoxy pyrimidines resulted in higher affinity aptamers from the 2’FY

selection. From this it was speculated that the increased affinity of 2’FY modified RNA is due

to formation of thermodynamically stable helices and rigid structures in 2’FY RNA compared

to the 2'amino modified RNA (12). However, X ray crystallography studies with 2’F modified

and unmodified duplexes demonstrated that the presence of 2’F modification does not alter the

overall helical structure of the duplex (21). The increased thermodynamic stability of the

helices is due to the highly electronegative fluorine polarizing the nucleobases, which enhances

the Watson-Crick pairing. Osmotic stress and X ray crystallography studies show that the

presence of 2’Fluorine also significantly reduces the hydration of the duplex, increasing its

hydrophobicity (22) (21). In our study also, we observed increased retention of 2’F poly U

sequences to the hydrophobic nitrocellulose membranes in the absence of target protein in

comparison to 2’OH RNA. Increased retention of nucleic acids to nitrocellulose membranes

was predominantly observed with oligonucleotides containing the poly U rich sequences

whereas past studies have reported the presence of Multi G Motifs (MGM’s) that exhibit

significant nonspecific nitrocellulose retention (23,24).

To localize the aptamer sequence in oligo 5183, we tested truncations, which identified

the 5’ poly U hairpin region, which was truncated to 39 nucleotides and tested for binding to

sGP by filter capture and EMSA assays and for the ability of a DNA oligo (5196)

Page 78: Selection of functional RNA aptamers against Ebola

70

complementary to the poly U loop reduce binding to sGP. These results confirmed that

oligonucleotide 5199 is an aptamer with high affinity for sGP and it was name 39SGP1A.

7\

and hydrogen bonding interactions among uracil (25) in comparison to GNRA (N any

nucleotide and R purine) loops or UUCG loops, which have interactions within loop bases(26).

39SGP1A contains an 18-nucleotide loop region comprised of 77% uracils with intermittent

purines and pyrimidines. The effect of 2’Fluoro modification on the structure of a loop is

unknown but, based on the reported effects of 2’FY substitution on helical stability due to base

polarization and increased hydrophobicity, we speculate that the high affinity of the 2’FY

modified 39SGP1A aptamer might be aided by strong intermolecular hydrogen bonding

interactions and π-π stacking interactions between the available 2’F uracils with sGP residues.

In addition, the 2’F- dependent structural organization within the loop is also expected to

contribute to enhanced affinity.

We investigated several of the selected oligonucleotides to to determine if they bind

the same epitope on sGP in a competition assay with 32P-39SGP1A and unlabeled low affinity

(oligo 4789), and high affinity (oligo 5179) and a DNA aptamer independently isolated

(manuscript in preparation). These three oligonucleotides competed with 39SGP1A for

binding sGP, which suggests that the protein has an epitope that is preferred for RNA binding.

39SGP1A was selected as a potential sensor for early EbolaVirus infection (27), which

would be performed with samples of patient blood. To obtain a significant signal to noise ratio

the aptamer should not bind other serum proteins. The most abundant serum protein is albumin

which is present in blood at concentrations as high as 300 μM. Although the selected

Page 79: Selection of functional RNA aptamers against Ebola

71

oligonucleotide, 5183, from which 39SGP1A was derived, bound to HSA with a Kd of 500

nM, 39SGP1A did not bind HSA. Thus, truncating the oligonucleotide to a minimal critical

sequence, appears to have eliminated structures that could bind HSA. 39SGP1A also did not

bind α2 macroglobulin and fibrinogen, two other serum proteins which could contribute to the

background noise.

In summary, we report the selection of a 2’FY-RNA aptamer that binds the Ebola virus

sGP glycoprotein with high affinity and specificity and demonstrate that the 2’F modification

is required for binding to sGP.

MATERIALS AND METHODS

SELEX library construction

A DNA single stranded oligonucleotide library named 487 was synthesized by IDT (Integrated

DNA Technologies, Coralville, IA) with the following sequence:

5’CCTGTTGTGAGCCTCCTGTCGAA (53N) TTGAGCGTTTATTCTTGTCTCCC 3’, and

N symbolizes equimolar mixture of A, C, G and T. The following primers were used for reverse

transcription and PCR reactions: Oligo484:5’-

TAATACGACTCACTATAGGGAGACAAGAATAAACGCTCAA-3’ and Oligo 485: 5’-

GCCTGTTGTGAGCCTCCTGTCGAA-3.

Page 80: Selection of functional RNA aptamers against Ebola

72

Conversion of ssDNA SELEX library to dsDNA

To generate starting RNA pool, an extension reaction was performed using primer 484

and the starting ssDNA pool. 10 reactions of 1000uls each containing 2 μM 487 pool, 3.3 μM

484 primer, 0.5 mM dNTP mix, 0.03 U/μl DNA Taq Polymerase (GenScript) in reaction buffer

containing 50 mM KCl, 10 mM Tris-HCl (pH 8.55), 1.5 mM MgCl2, 0.1% TritonX-100 was

incubated at 94°C for 5 min, 65°C for 15 min and 72°C for 99 min using the Multi GENE II

PCR minicycler. The generated dsDNA was run on 2% agarose gel and purified using Qiagen

PCR purification kit. Purified DNA was quantified via Nanodrop.

RNA synthesis and purification

RNA was prepared by in-vitro transcription using DurascribeTM T7 transcription kit

from (Epicentre, Madison). Around 2nmoles of dsDNA from pool generated from extension

is incubated with 5 mM ATP, 5 mM GTP,5 mM 2’FY UTP,5 mM 2’FYCTP, 5 mM DTT,

0.2U/μL T7 Enzyme in a total volume of 2.8 mL and 5% DMSO and incubated at 37°C for 4

h. DNA mixture was then digested with 1 MBU DNAse I. The RNA was resolved through a

7M urea 8% (19:1 acrylamide:bisacrylamide) gel in 1X Tris Borate EDTA buffer, pH 9.1 to

separate the residual NTPs and abortive transcripts. Transcribed RNA was eluted in TE buffer

(10 mM Tris-HCl, 1 mM EDTA pH 8.0) for 18 h at 37°C.

In-vitro selection of RNA aptamers against sGP (soluble glycoprotein)

In vitro selection of an aptamer for sGP was carried out by SELEX (Systematic

Evolution of Ligands by EXponential enrichment). Prior to the first round of selection, a

Page 81: Selection of functional RNA aptamers against Ebola

73

dsDNA pool was generated from a library of ssDNAs (487D pool) by primer extension.

Starting RNA pool was generated by in-vitro transcription using dsDNA (487D) pool as

template. RNA pool (1X1015 molecules) 2nmoles was incubated with sGP and the non-binding

RNA sequences were partitioned by passing it through nitrocellulose membrane. For each

positive selection, to increase the stringency of selection ratio of RNA pool: sGP was

increased. To remove nonspecific membrane binders negative selections were performed

against the nitrocellulose membrane. Reverse transcribed polymerase chain reaction (RT-PCR)

amplification was used to generate DNA for subsequent selection rounds. 8 rounds of

selections were performed including negative and positive selections. Enriched pool from the

10th SELEX round was cloned into the TOPO XL PCR cloning plasmid for sequencing, in

addition PCR libraries were prepared for Next Gen Sequencing using Illumina sequencing

primers.

Nitrocellulose capture assay

To determine the binding affinity of sGP to RNA aptamers, 5’P32 end labeled RNA

was firstly prepared. 2’FY modified RNA was transcribed using Durascribe transcription kit

(Epicentre Technologies), transcribed RNA was electrophoresed in an 8% polyacrylamide–7

M urea gel, and eluted from the gel. RNA was dephosphorylated with calf intestine

phosphatase for 1 h at 37°C, extracted with phenol, and precipitated with ethanol. The RNA

was end-labeled with [γ-32P] ATP and T4 polynucleotide kinase. To determine the binding

affinity of the RNA aptamer 5183 to sGP, binding studies were performed by first denaturing

the 2nM RNA at 95 degrees for 5 mins followed by refolding in binding buffer [137 mM NaCl,

Page 82: Selection of functional RNA aptamers against Ebola

74

2.7 mM KCl, 10 mM Na2HPO4, 2 mM KH2PO4, 5 mM MgCl2, pH 7.4] at RT for 30 min. RNA

was incubated with varying concentrations of sGP (10 nM-1 μM) at RT for 30 min. The RNA

sGP complex was passed through nitrocellulose membranes (HAWP 02500) and filters were

washed with 3 mL binding buffer and later quantified using Liquid Scintillation Counter. The

data was fit to F=Fmin+(Fmax*L^n)/(L^n+Kd^n) to determine the Kd.

Electrophoretic mobility shift assay (EMSA)

Binding reactions performed by firstly by denaturing of 32P- 5’ end labeled RNA at 95

degrees for 5mins followed by slow cooling in binding buffer [137 mM NaCl, 2.7 mM KCl,

10 mM Na2HPO4, 2 mM KH2PO4, 5 mM MgCl2, pH 7.4] for 30 min. Refolded RNA was

incubated with sGP in binding buffer for 30 min at room temperature. The mixtures were

analyzed for complex formation by resolution through a 6% native polyacrylamide gels run in

Tris HEPES EDTA buffer (33 mM Tris, 66 mM HEPES, 0.1 mM EDTA-Na, pH 7.5) at 200V

and the 32P was visualized using a phosphor imager.

Competition binding assay

Competitive binding reactions were performed by incubating 4 μM sGP with 18 μM

competitor RNA sequences for 30 min at RT in binding buffer. Following the preincubation

period, 32P-labeled 39sGP1A 2’FY-RNA was added to the reaction mix to a final concentration

of 200 nM and the samples were incubated for 15 min. The RNA protein complexes were then

UV crosslinked at 365 nm for 15 min. Crosslinked products were run on 10% SDS PAGE

reducing gel, the gel was dried and later imaged by Typhoon scanner and quantification was

Page 83: Selection of functional RNA aptamers against Ebola

75

performed using Image J software. 39sGP1A RNA aptamer was tested individually against

39sGP1A, 5179, 4789 and 6011 oligonucleotides.

Incorporation of 4 Thio UTP and RNA protein UV crosslinking

RNA was invitro transcribed in the presence of 2’F-UTP and 4-thio-uridine

5’triphosphate (Trilink Biotechnologies) at a molar ratio of 1:1) and radiolabeled with P32-γ-

ATP by T4 polynucleotide kinase. 3 pmol of RNA was denatured at 95°C for 5 min followed

by snap cooling on ice for 5 min. The RNA was folded in PS Buffer (137 mM NaCl, 2.7 mM

KCl, 10 mM Na2HPO4, 2 mM KH2PO4, 5 mM MgCl2, pH 7.4) for 10 min at RT. To link sGP

to the 2’FY-RNA aptamer 5183, 3 pmol 2’FY-RNA was incubated with 15 mol sGP in 15 μL

binding buffer for 20 min at RT for complex formation. This mixture was irradiated for 6 min

and 12 min under a UV lamp (365 nm) with the samples placed 3 cm from the lamps. RNA-

sGP samples were treated with proteinase K for 30 min. The samples were later resolved

through a 10% reducing SDS PAGE gel to separate the 2’FY-RNA-protein complex, free

2’FY-RNA and free protein. The 32P on the gel was visualized using a phosphor imager.

Antisense competition assay with DNA oligo

39sGP1A RNA aptamer was transcribed using Durascribe T7 transcription kit, RNA

transcript was dephosphorylated using alkaline phosphatase (Promega 1U) at 37°C for 1 h,

followed by phenol chloroform extraction. The dephosphorylated 2’FY-RNA was 5’ end

labeled with 32P by incubating with T4 polynucleotide kinase (NEB) at 37°C for 30 min

followed by 65°C for 15 min to inactivate the enzyme. The labeled RNA aptamer was

Page 84: Selection of functional RNA aptamers against Ebola

76

denatured at 95°C for 5 min followed by snap cooling on ice for 5 min and then incubated at

RT for 10 min in (137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 2 mM KH2PO4, 5 mM

MgCl2, pH 7.4). The folded RNA was incubated in the presence and absence of 5-fold excess

of oligonucleotide 5196 for 20 min then sGP was added and the samples incubated for 30 min.

The samples were then irradiated for 15 min UV (365nm) at 3 cm distance from the lamp. The

samples were later resolved through a 10% reducing SDS PAGE gel to separate the 2’FY-

RNA aptamer-protein complex, free RNA and free protein. The 32P on the gel was visualized

using phosphor imager.

ACKNOWLEDGEMENT

We thank Lee Bendickson and Muslum Ilgu for helpful discussions and advice. This

work was funded by NIH grant 5R43AI118139-02 to Aptalogic Inc.

Page 85: Selection of functional RNA aptamers against Ebola

77

LEGENDS TO FIGURES

Figure 1. Next Gen Sequencing Analysis and Consensus Motif Identification. A) The

frequency of singletons, enriched and unique sequences for each SELEX round B) The SELEX

protocol for selecting 2’FY-RNA aptamers with high affinity for sGP C, D) The frequency of

base occurrence in each position for the 1st and 9th round pools showing evidence of sequence

enrichment during selection E) A conserved sequence motif is identified among the top

identified clusters.

Figure 2. The Affinities of Selected 2’FY-oligonucleotides for sGP. A) The concentration

dependence of sGP binding to 2 nM 32P-5183 as a 2’FY-RNA and an RNA oligonucleotide.

The estimated Kd was for the 2’FY-RNA was 50 nM. B) An extended sGP titration up to 3

μM against 2nM nM oligonucleotide 32P-5183 to validate saturation of the binding curve, C)

The concentration dependence of sGP binding to 20 nM

32P- oligonucleotide 4789. The estimated Kd was 500 nM. D) M fold predicted secondary

structures of oligonucleotide s 4789 and 5183 E) Electrophoretic mobility shift assay with 100

nM 32P- oligonucleotide 5183. Buff, buffer only. sGP, buffer plus 2 μM sGP. comp, buffer

plus 2 μM sGP and 10 μM Lcn2 aptamer.

Figure 3. Determining the Binding Epitope for sGP on the 2’FY RNA aptamer.

A) A mutation count chart created in APTAGUI to determine the mutation frequency for 5183

at each nucleotide position, B) The M-fold predicted secondary structures of truncations of

5183, C) UV crosslinking of 200 nM 32P 39SGP1A to 2.5 uM sGP followed by treatment with

or without proteinase K, D) Binding of 10nM 32P-oligo 5183 truncated oligonucleotides to

50nM and 500nM sGP, E) Binding of 0.5 uM 32P-39SGP1A to 10uM sGP in the presence of

Page 86: Selection of functional RNA aptamers against Ebola

78

5 uM antisense oligonucleotide 5196, which is complimentary to the loop region of

oligonucleotide 39SGP1A, F) Concentration dependence of sGP binding to 10nM 32P-

oligonucleotide 39SGP1A and 10 nM 32P- oligonucleotide 5015. The estimated Kd was 27nM

for 39SGP1A

Figure 4. Competition Binding Assay to Determine if the Selected Aptamers Bind to the

Same Protein Epitope. 32P- oligonucleotide 39SGP1A (0.2 μM) was incubated for 15 min at

23°C with a 10-fold excess of the identified oligonucleotides. The samples were stabilized by

UV crosslinking and resolved by SDS-PAGE.

Figure 5. Specificity of 2’FY-5183 and 39SGP1A for sGP. A) Binding of 2 nM 2

’FY 5183 protoaptamer to human serum albumin (HSA) B) Binding of the 10nM 32P 2’FY

39SGP1A aptamer to HSA C) Lack of affinity of 10 nM 2’FY 39SGP1A for human serum

albumin (200 nM to 50 μM), fibrinogen (200 nM to 4 μM) or α2macroglobulin (200 nM to 5

μM).

Page 87: Selection of functional RNA aptamers against Ebola

79

REFERENCES

1. Sanchez, A., Trappier, S. G., Mahy, B., Peters, C. J., and Nichol, S. T. (1996)

The virion glycoproteins of Ebola viruses are encoded in two reading frames and are expressed

through transcriptional editing. Proceedings of the National Academy of Sciences 93, 3602-

3607

2. Volchkov, V. E., Feldmann, H., Volchkova, V. A., and Klenk, H.-D. (1998)

Processing of the Ebola virus glycoprotein by the proprotein convertase furin. Proceedings of

the National Academy of Sciences 95, 5762-5767

3. Pallesen, J., Murin, C. D., de Val, N., Cottrell, C. A., Hastie, K. M., Turner, H.

L., Fusco, M. L., Flyak, A. I., Zeitlin, L., and Crowe Jr, J. E. (2016) Structures of Ebola virus

GP and sGP in complex with therapeutic antibodies. Nature microbiology 1, 16128

4. de La Vega, M.-A., Wong, G., Kobinger, G. P., and Qiu, X. (2015) The multiple

roles of sGP in Ebola pathogenesis. Viral immunology 28, 3-9

5. Tuerk, C., and Gold, L. (1990) Systematic evolution of ligands by exponential

enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 249, 505-510

6. Ellington, A. D., and Szostak, J. W. (1990) In vitro selection of RNA molecules

that bind specific ligands. Nature 346, 818-822

7. Robertson, D. L., and Joyce, G. F. (1990) Selection in vitro of an RNA enzyme

that specifically cleaves single-stranded DNA. Nature 344, 467-468

8. Auweter, S., and Allain, F. H.-T. (2008) Structure-function relationships of the

polypyrimidine tract binding protein. Cellular and Molecular Life Sciences 65, 516-527

Page 88: Selection of functional RNA aptamers against Ebola

80

9. Pérez, I., Lin, C.-H., McAFEE, J. G., and Patton, J. G. (1997) Mutation of PTB

binding sites causes misregulation of alternative 3'splice site selection in vivo. Rna 3, 764-778

10. Singh, R., Valcarcel, J., and Green, M. R. (1995) Distinct binding specificities

and functions of higher eukaryotic polypyrimidine tract-binding proteins. Science 268, 1173

11. Pieken, W. A., Olsen, D. B., Benseler, F., Aurup, H., and Eckstein, F. (1991)

Kinetic characterization of ribonuclease-resistant 2'-modified hammerhead ribozymes. Science

253, 314-318

12. Pagratis, N. C., Bell, C., Chang, Y.-F., Jennings, S., Fitzwater, T., Jellinek, D.,

and Dang, C. (1997) Potent 2′-amino-, and 2′-fluoro-2′-deoxyribonucleotide RNA inhibitors

of keratinocyte growth factor. Nature biotechnology 15, 68-73

13. Levine, H. A., and Nilsen-Hamilton, M. (2007) A mathematical analysis of

SELEX. Computational biology and chemistry 31, 11-35

14. Hoinka, J., Dao, P., and Przytycka, T. M. (2015) AptaGUI—A graphical user

interface for the efficient analysis of HT-SELEX data. Molecular therapy. Nucleic acids 4,

e257

15. Bailey, T. L., Boden, M., Buske, F. A., Frith, M., Grant, C. E., Clementi, L.,

Ren, J., Li, W. W., and Noble, W. S. (2009) MEME SUITE: tools for motif discovery and

searching. Nucleic acids research 37, W202-W208

16. Zuker, M., Mathews, D. H., and Turner, D. H. (1999) Algorithms and

thermodynamics for RNA secondary structure prediction: a practical guide. in RNA

biochemistry and biotechnology, Springer. pp 11-43

Page 89: Selection of functional RNA aptamers against Ebola

81

17. Zhai, L., Wang, T., Kang, K., Zhao, Y., Shrotriya, P., and Nilsen-Hamilton, M.

(2012) An RNA aptamer-based microcantilever sensor to detect the inflammatory marker,

mouse lipocalin-2. Anal Chem 84, 8763-8770

18. Bhaskar, V., Roudko, V., Basquin, J., Sharma, K., Urlaub, H., Séraphin, B., and

Conti, E. (2013) Structure and RNA-binding properties of the Not1–Not2–Not5 module of the

yeast Ccr4–Not complex. Nature structural & molecular biology 20, 1281-1288

19. Luo, G. (1999) Cellular proteins bind to the poly (U) tract of the 3′ untranslated

region of hepatitis C virus RNA genome. Virology 256, 105-118

20. Kanai, A., Tanabe, K., and Kohara, M. (1995) Poly (U) binding activity of

hepatitis C virus NS3 protein, a putative RNA helicase. FEBS letters 376, 221-224

21. Pallan, P. S., Greene, E. M., Jicman, P. A., Pandey, R. K., Manoharan, M.,

Rozners, E., and Egli, M. (2010) Unexpected origins of the enhanced pairing affinity of 2′-

fluoro-modified RNA. Nucleic acids research, gkq1270

22. Patra, A., Paolillo, M., Charisse, K., Manoharan, M., Rozners, E., and Egli, M.

(2012) 2′‐Fluoro RNA Shows Increased Watson–Crick H‐Bonding Strength and Stacking

Relative to RNA: Evidence from NMR and Thermodynamic Data. Angewandte Chemie

International Edition 51, 11863-11866

23. Shi, H., Fan, X., Ni, Z., and Lis, J. T. (2002) Evolutionary dynamics and

population control during in vitro selection and amplification with multiple targets. RNA 8,

1461-1470

Page 90: Selection of functional RNA aptamers against Ebola

82

24. Tuerk, C., MacDougal, S., and Gold, L. (1992) RNA pseudoknots that inhibit

human immunodeficiency virus type 1 reverse transcriptase. Proceedings of the National

Academy of Sciences 89, 6988-6992

25. Sashital, D. G., Venditti, V., Angers, C. G., Cornilescu, G., and Butcher, S. E.

(2007) Structure and thermodynamics of a conserved U2 snRNA domain from yeast and

human. Rna 13, 328-338

26. Sheehy, J. P., Davis, A. R., and Znosko, B. M. (2010) Thermodynamic

characterization of naturally occurring RNA tetraloops. RNA

27. Bruno, J. G. (2015) Predicting the Uncertain Future of Aptamer-Based

Diagnostics and Therapeutics. Molecules 20, 6866-6887

Page 91: Selection of functional RNA aptamers against Ebola

83

Figure 1

Next Gen Sequencing Analysis And Consensus Motif Identification

Page 92: Selection of functional RNA aptamers against Ebola

84

Figure 2

The Affinities of Selected 2’FY-oligonucleotides for sGP.

Page 93: Selection of functional RNA aptamers against Ebola

85

Figure 3

Determining the Binding Epitope for sGP on the 2’FY RNA Aptamer.

Page 94: Selection of functional RNA aptamers against Ebola

86

Figure 4

Competition Binding Assay to Determine if the Selected Aptamers Share a

Protein Epitope

Page 95: Selection of functional RNA aptamers against Ebola

87

Figure 5

Specificity of 2’FY-5183 and 39SGP1A for sGP.

Page 96: Selection of functional RNA aptamers against Ebola

88

CHAPTER 3

SELECTING A FUNCTIONAL ANTI–VIRAL RNA APTAMER AGAINST

EBOLAVIRUS SURFACE GLYCOPROTEIN

Shambhavi Shubhama,b, Jan Hoinkac, Emma Swansona, Teresa M. Przytyckac,

Wendy Mauryd, and Marit Nilsen-Hamiltona,b

aIowa State University, Ames IA, bAptalogic Inc., Ames IA, cNational Institutes of Health,

Bethesda, DC, dUniversity of Iowa, Iowa City IA

Experimental contributions to manuscript:

SELEX and identification of potential sequences and binding assays to identify oligonucleotide

sequences that bind to GP1,2 Δmucin was done by Shambhavi Shubham.

Analysis of Next Gen Sequencing data was performed by Jan Hoinka alongwith the

identification of top ten clusters.

Emma Swanson performed the binding assays for oligonucleotide sequences that bind to GP1,2

full length.

Page 97: Selection of functional RNA aptamers against Ebola

89

ABSTRACT

Ebola viruses (EBOV) cause severe disease in humans and in non-human primates in the form

of viral hemorrhagic fever. Ebola virus is also of public health concern as a potential

bioterrorism organism for which no vaccine or anti-viral is available. Fast-acting and

prophylactic therapeutics are needed to reduce mortality. In view of the paucity of current

antiviral therapies for EBOV we undertook to select an aptamer against the EBOV surface-

exposed glycoprotein, GP1. Here we described the selection and characterization of several

2’FY RNA aptamers that recognize the GP1,2 heterotrimer. One aptamer recognizes the

heterotrimer with the mucin domain included, which is the form of the protein found on the

circulating virion. Another aptamer recognizes the GP1,2 heterotrimer lacking the mucin

domain of GP1, which is the form of the protein in the endosomes that mediates entry into the

cytoplasm. Thus, these aptamers target the virion in two stages of infection and together might

be effective in preventing the interaction between virus and the host cell and viral entry into

the cytoplasm, thereby reducing viremia and spread of the virus to others.

Page 98: Selection of functional RNA aptamers against Ebola

90

INTRODUCTION

Ebola Virus (EBOV) and Marburg Virus (MARV) are members of the family Filoviridae of

enveloped viruses with a non-segmented negative strand RNA genome. Filoviruses cause

sporadic outbreaks of hemorrhagic fever in human and non-human primates in Africa with

fatality rates up to 90%. Filovirus hemorrhagic fever is associated with high levels of

inflammatory cytokines and coagulation disorders resulting in septic shock and multiorgan

failure (1). The virus is transmitted through contact with bodily fluids and can infect various

cell types across different host species. They typically infect hepatocytes, dendritic cells,

endothelial cells, macrophages, and monocytes (2). The viral genome encodes seven structural

proteins: envelope glycoprotein (GP), major matrix protein (VP40), nucleoprotein (NP),

polymerase cofactor (VP35), replication/transcription protein (VP30), minor matrix protein

(VP24), and RNA dependent DNA polymerase (L) (3). The viral replication cycle involves

attachment of the virus to its receptors, uptake of the virus, intracellular trafficking of the virus

in the endosome and release of the nucleocapsid into the cytoplasm, synthesis of viral proteins,

assembly and budding from the cell surface (4).

Viral glycoprotein (GP1) is a homotrimer, forming spikes on the surface of the virus and is the

only protein on the EBOV viral particle surface. Folding and assembly of GP1 occurs

independently of other viral proteins (5). It is initially synthesized as a precursor that is later

cleaved by the proprotein convertase, furin, into GP1 (140kD) and GP2 (26kD) that are linked

by disulfide bonds (6). The GP2 subunit contains two heptad repeat regions, which facilitate

assembly of GP into trimers, a transmembrane anchor sequence, and the fusion loop (5). The

GP1 subunit contains the receptor binding site and a heavily glycosylated mucin domain. The

Page 99: Selection of functional RNA aptamers against Ebola

91

residues responsible for mediating interactions between GP1 and the host cell are in the 230-

residue N-terminal domain of GP1. The crystal structure of EBOV GP demonstrates that the

receptor binding site of GP1 is masked by the glycan cap created by the mucin domain (7).

Consistent with this structure is the observation that deletion of the C-terminal mucin domain

enhances viral infectivity in vitro (8-10).

Ebola viruses infect a broad range of cells and a host of proteins are known to enhance viral

entry into the host cells including the C-type lectins L-SIGN, DC-SIGN and the tyrosine kinase

receptor Axl (11-14). Infectivity involves at least two recognition events 1) endocytosis at the

host cell surface and 2) cytoplasmic entry in the endosome. The cell surface receptor for EBOV

is believed to be the T-cell immunoglobulin and mucin domain 1 (TIM-1) receptor (15). The

usual function of the TIM-1 receptor is to bind phosphatidyl serine on the surface of apoptotic

cells and facilitate their phagocytosis (16). TIM-1 also facilitates the entry of Dengue Virus by

directly interacting with the virion-associated phosphatidylserine (17). Structural studies of the

TIM immunoglobulin domain have demonstrated that phosphatidyl serine binds in a cavity

built up by the CC’ and FG loops termed the metal ion-dependent ligand binding site (MILIBS)

(17). By contrast with the Dengue virus interaction, TIM-1-mediated Ebola infection depends

on a direct interaction between the viral glycoprotein GP through residues in the ARD5

domain, which is outside of the TIM-1 phosphatidylserine binding pocket and does not include

the MILIBS (15). This specificity was demonstrated by showing that the anti-TIM antibody,

which binds the ARD5 epitope, completely blocked EBOV entry to Vero cells, whereas A6G2

antibody, which prevents TIM-1 binding to phosphatidyl serine, was much less effective in

blocking the viral transduction (15). Co-immunoprecipitation studies also showed that there is

Page 100: Selection of functional RNA aptamers against Ebola

92

direct interaction between Ebola GP and TIM-1 that depends on the receptor binding domain

(15). However, later it was discovered that phosphatidyl serine on the virus surface directly

binds TIM-1 thereby promoting virion internalization (18).

The endosomal receptor for EBOV is believed to be Niemann-Pick C1(NPC1), a lysosomal

cholesterol transporter (19). The binding site for C1(NPC1) is exposed after endosomal

cysteine proteases cleave EBOV GP1 to remove the heavily glycosylated C terminal region to

generate an entry intermediate comprising of the N terminal of GP1 and GP2 (19,20).

Mutations in the C1(NPC1) binding site on GP1 decrease viral entry (19).

In this study, we report the selection and characterization of 2’FY RNA aptamers against

GP1,2 delta mucin and full length GP1,2 glycoproteins. Together these aptamers are expected

to thwart interactions between virion and host at two stages. The GP1,2 delta mucin

conformation with exposed RBD interacts with NPC-1 in the endosome. Therefore, an RNA

aptamer against the GP1,2 delta mucin could intercept the viral host interaction at that step. In

addition, TIM-1 is not expressed by all EBOV permissive cell lines ,(18), which suggests the

occurrence of other unidentified cellular receptors for EBOV. In this situation, an RNA

aptamer against of GP1,2 full length will be effective in preventing viral replication.

RESULTS

SELEX and Analysis

To identify aptamers that bind Ebola GP1,2 we performed two SELEX experiments in which

early rounds selected against the purified recombinant protein and later rounds were against

the protein in the context of the viral particle (Fig. 1). GP1,2 lacking the mucin domain

Page 101: Selection of functional RNA aptamers against Ebola

93

(GP1,2Δmucin) was used in the early rounds of SELEX 1280 and the full length GP1,2 was

used in the early rounds of SELEX 1281 (Fig. 1A,D). In both protocols, later rounds were

selected against Vesicular Stomatitis Virus (VSV) pseudotyped with GP1,2Δmucin

(SELEX1280) or GP1,2 (SELEX1281).Next Gen Sequencing results from both protocols

showed evidence of expanding potential aptamer populations with a reduction in unique

sequences from early to late rounds, which indicates population expansions (Fig. 1B, E). The

progress of oligonucleotide selection evaluated from frequency plots for each round in which

the number (counts) of each base (Y axis) is plotted for each base position in the

oligonucleotide length (X axis). The frequency plots show that a different population emerged

in SELEX1280 after switching the selection from the protein to the virus particles, whereas

this did not happen in SELEX1281 (Fig. 1C, F). We selected the top 10 enriched classes from

each SELEX experiment for further study.

From the SELEX experiments we identified clusters of selected sequences families using the

APTAGUI web application. To further identify which clusters likely contained high affinity

aptamers, oligonucleotides representing the top 10 sequence clusters found for SELEX 1280

were mixed in equal amounts and incubated for 15 min with VSV alone or VSV pseudotyped

with GP1,2Δmucin (RNA/Protein =10:1). The samples were then spun through a sucrose

cushion to separate the RNAs bound to the virus particles from the unbound RNAs. The

collected RNAs were cloned and sequenced. Oligonucleotides 4789 and 4796 were the most

highly represented of oligonucleotides that remained associated with the virus particles through

the sucrose cushion, summing to 47% of the clones from the GP1,2Δmucin-VSV population

(Fig. 2). These oligonucleotides also had the highest selective binding to the GP1,2Δmucin-

Page 102: Selection of functional RNA aptamers against Ebola

94

VSV captured population, being represented 3-4.5 times as much with GP1,2Δmucin-VSV

than with the VSV control particles.

The Dissociation Constant (Kd) of Selected Proto-Aptamers

Oligonucleotides 4789 and 4796 (SELEX 1280), identified from the sucrose centrifugation

assay, were tested for their affinities for the GP1,2 delta mucin recombinant and GP1,2 full

length recombinant proteins using the nitrocellulose filter capture assay. Oligonucleotide 4789

bound GP1,2Δmucin with a Kd of 140nM but did not bind the full length GP1,2 glycoprotein

(Fig. 3A). These results suggest that 4789 binds to an epitope on GP1,2 that is inaccessible in

the presence of mucin domain. The interaction between oligonucleotide 4789 and GP1,2 was

also demonstrated by EMSA in which a shifted band was observed when the RNA was

incubated with GP1,2Δmucin. The specificity of oligonucleotide 4789 binding for

GP1,2Δmucin was also testing with BSA for which no binding was observed. (Fig. 3B).

4797 oligonucleotide sequence appeared in the top clusters of both SELEX 1280 and 1281

and was tested for its binding to GP1,2 full length and GP1,2 delta mucin recombinant protein.

Interestingly, it bound neither protein. This sequence appeared in the top clusters, which were

performed over different time periods. This might be an example of the evolution of a “sticky”

oligonucleotide that binds nonspecifically to many surfaces such as the nitrocellulose

membrane used for particle capture. The “sticky” nature of this oligonucleotide is also

consistent with the observation that 4797 pelleted predominantly with VSV control particle

when passed through a sucrose gradient.

Page 103: Selection of functional RNA aptamers against Ebola

95

Oligonucleotides 5185, 5186, 5187 and 5188 sequences selected in SELEX 1281 in but not in

SELEX 1280 were tested for binding to the GP1,2 full length protein. The affinities (Kd) of

these oligonucleotides were 50 nM, 120 nM, 560 nM and 300 nM for 5185, 5186, 5187 and

5188 respectively. These oligonucleotides tested for binding to the isolated GP1,2 proteins

were further evaluated for their ability to bind virus particles pseudotyped with full-length

GP1,2 and with control virus particles lacking GP1,2. Oligonucleotide 5185 bound GP1,2 full

length glycoprotein with a Kd of 50 nM and it also bound the GP1,2 pseudotyped virus

particles in a titration that achieved saturation. (Fig.4).

DISCUSSION

Aptamers have notable advantages over antibodies that include non-immunogenicity, in-vitro

production and stability (21).A variety of aptamers have been selected to counter host-virus

interactions by targeting different viral proteins (21-23). Examples include: 1) An RNA

aptamer that recognizes HIV-GP120 that effectively neutralizes clinical isolates of HIV (24),

2) Viral RNA-dependent RNA polymerases with basic patches that permit easier selection of

an RNA aptamer (25), and 3) the internal ribosome entry site (IRES), a structured RNA region

in the viral mRNA that binds the ribosome and initiates cap-independent translation to favors

viral protein translation. Aptamers targeting the IRES inhibit IRES-dependent translation of

HCV proteins (26). These cited aptamers were selected to counter host–virus interactions by

targeting viral components.

In this study 2’FY modified RNA aptamers were identified that bind to two conformations of

Ebola glycoproteins which could be used to thwart viral host cell interactions. SELEX 1280

Page 104: Selection of functional RNA aptamers against Ebola

96

was performed against GP1,2 Δmucin and SELEX1281 was performed against full length

GP1,2. We employed Next Gen Sequencing (NGS) of the random region of several SELEX

populations to identify expanding potential aptamer populations. NGS has the advantage over

conventional sequencing that it provides up to 130 million sequence reads and, with barcoding,

multiple rounds can be sequenced simultaneously. In these SELEX experiments, reduction in

unique sequences from 87% to 23% of the total between the 2nd and 5th rounds demonstrated

selection for a subset of sequences. The 9th round involved selection against virus particles and

the frequency plot showed that a different population emerged in this round for SELEX 1280.

This change in dominant populations may be due to the elimination of oligonucleotides that

bound to surfaces on the viral protein other than the protein surface exposed on the virus

particles. In contrast, there wasn’t much increase in the percentage of unique sequences in

round 9. The unique oligonucleotide sequences are expected to be oligonucleotides that bind

weakly to the protein or that do not bind and that were therefore retained through the washing

during filter capture.

For SELEX 1281, 1st, 4th, 5th, 8th and 10th rounds were barcoded by PCR and sequenced. 27

million reads were obtained with same number of sequences obtained from each round. The

uniqueness declined from each round, with the first round having 99.6% in the unique fraction

and the percent of unique sequences reduced with each cycle. The last round of selection had

25.9% unique sequences. Reduction in the unique fraction in both SELEX experiments clearly

suggests that sequences were enriched as we moved along in the selection process. In contrast

to SELEX1281, in which there was change in the frequency plot when the target was changed

Page 105: Selection of functional RNA aptamers against Ebola

97

from protein to virus particle expressing protein, no change in population was observed when

the target was changed from recombinant GP1,2 to virus particles expressing GP1,2.

Our overall goal was to obtain RNA aptamers that can be applied in the blood in which they

must remain associated with the viral particles in the presence of shear stress. Therefore,

selection was performed under conditions that apply sheer to the aptamer-viral complexes. The

RNAs bound to viral particles were spun through a sucrose cushion to identify RNAs retained

on the virus particles after subjection to shear stress. For this assay the top ten potential

aptamers identified from SELEX 1280 were mixed at concentrations to achieve high

RNA/Protein (10:1). Centrifugation was for an hour to select for aptamers that form complexes

with slower koffs (27). The kinetic half-lives of aptamer protein complexes with pM Kd’s

range from 6 min to 60 mins.

Using this method, we selected for oligonucleotides with the sequences 4789 and 4796 that

preferentially bound to the GP1,2Δmucin pseudotyped VSV virus particles. Oligonucleotide

4789 bound to GP1,2 Δmucin with a Kd of 143 nM and it showed concentration dependent

saturation binding against GP1,2Δmucin virus particles. We did not observe any binding with

the VSV control virus particles. From the top clusters obtained from SELEX 1281, we tested

the binding of oligonucleotides 5185, 5186, 5187 and 5188 to full length GP1,2. All the

oligonucleotides bound to full length GP1,2 with different affinities. Among these,

oligonucleotide 5185 bound full length GP1,2 with high affinity (50 nM) Kd and also bound

the GP1,2 pseudotyped virus particles. Thus, from the binding assay and sucrose centrifugation

assay we can conclude that oligonucleotides 4789 and 5185 are sequences that bind to

Page 106: Selection of functional RNA aptamers against Ebola

98

GP1,2Δmucin and full length GP1,2 conformations, in recombinant form and when expressed

on virus particles.

GP1,2 interacts with the NPC-1 receptor in endosome in low pH conditions therefore the next

step in the development of these aptamers will be to test the sensitivity of the aptamer-GP1,2

interaction at low pH. The identified oligonucleotides that bind full-length GP1,2 pseudotyped

virus particles will be tested to determine if they prevent viral replication using an in vitro

transduction assay. Once an aptamer that reduces viral infectivity in vitro has been identified,

the efficacy of these aptamers against the viruses will be determined in an in vivo assay. If

these identified oligonucleotides that bind to the GP1,2 proteins do not prevent viral

transduction they could still be used as drug and anti-viral agent carriers to tag along with the

virus particles and target the cells that are also targeted by the virus (28,29). They could also

be developed for diagnostic purposes (21,30). For EBOV, a rapid diagnosis of infection is

required because, after the appearance of symptoms, the patient succumbs to infection in 10

days. The current methods for diagnosis of EBOV are RT-PCR and ELISA (31). But these are

time consuming assays that require sophisticated and expensive equipment. The necessary

equipment for these assays is generally lacking in regions of EBOV occurrence. Aptamers are

compatible with inexpensive analytical equipment that could be readily used in field to detect

EBOV and other analytes.

Page 107: Selection of functional RNA aptamers against Ebola

99

MATERIALS AND METHODS

SELEX library construction

DNA single stranded oligonucleotide library named 487 was generated

5’CCTGTTGTGAGCCTCCTGTCGAA (53N) TTGAGCGTTTATTCTTGTCTCCC 3’, and

N symbolizes an equimolar mixture of A, C, G and T. It was synthesized by Integrated DNA

Technologies, Coralville, IA. Primers given below were used for reverse transcription and PCR

reactions.

484:5’TAATACGACTCACTATAGGGAGACAAGAATAAACGCTCAA-3’ and

485: 5’-GCCTGTTGTGAGCCTCCTGTCGAA-3.

Conversion of ssDNA SELEX library to dsDNA

To generate a starting RNA pool, an extension reaction was performed using primer

484 and the starting ssDNA pool. 10 reactions of 1000 μL each containing 2uM 487 pool,

3.3uM 484 primer, 0.5 mM dNTP mix, 0.03 U/μL DNA Taq Polymerase (GenScript) in

reaction buffer containing 50 mM KCl, 10 mM Tris-HCl (pH 8.55), 1.5 mM MgCl2, 0.1%

TritonX-100 was incubated at 94°C for 5 min, 65°C for 15 min and 72°C for 99 min using the

Multi GENE II PCR minicycler. The generated dsDNA was run on 2% agarose gel and purified

using Qiagen PCR purification kit. Purified DNA was quantified using a Nanodrop

spectrophotometer.

Page 108: Selection of functional RNA aptamers against Ebola

100

RNA Synthesis and Purification

RNA was prepared by in-vitro transcription using a DurascribeTM T7 transcription kit

(Epicentre, Madison). Around 2 nmole of dsDNA from a pool generated from extension was

incubated with 5 mM ATP, 5 mM GTP, 5 mM 2’FY UTP, 5 mM 2’FY CTP, 5 mM DTT, 0.2

U/μL of T7 polymerase in a total volume of 2.8 mL and 5% DMSO and incubated at 37°C for

4 h. The DNA was then digested with 1MBU of DNAse I and the remaining RNA was resolved

through an 8% (19:1 acrylamide:bisacrylamide) acrylamide gel in 7M urea, Tris Borate EDTA

buffer (TBE buffer 89 mM Tris, 89 mM Boric acid and 2 mM EDTA) pH 9.1 to separate

residual NTPs and abortive transcripts. The transcribed RNA was eluted in 10 mM Tris-HCl,

1 mM EDTA pH 8.0 for 18 h at 37°C and concentrated by ethanol precipitation.

In-Vitro Selection of RNA Aptamers That Bind Ebola Virus Particles and Glycoproteins

SELEX was performed against GP1,2 full length glycoprotein SELEX 1281 and GP1,2

mucin domain deleted glycoprotein (SELEX1280). A ssDNA pool with a complexity of 453

sequences from IDT Technologies was used as starting pool. This pool was designated as 487,

with each draw from the tube identifying the subset of oligonucleotides with a different letter

(487A and 487D were used here). Constant regions that flank the 53-base central region of

random sequence in pool 487 are complementary to primers 484 and 485. In the first round of

SELEX, DNA polymerase catalyzed extension copied the ssDNA to produce dsDNA from the

forward primer TAATACGACTCACTATAGGGAGACAAGAATAAACGCTCAA (484),

which contains a T7 promoter. The DNA was in-vitro transcribed using T7 RNA polymerase,

2’ fluoro-labeled pyrimidine triphosphates, and unmodified purine triphosphates. The resulting

Page 109: Selection of functional RNA aptamers against Ebola

101

complexity was about 1015 RNA sequences. The RNA pool was incubated with the GP1 full

length and mucin domain deleted trimer at a 1:1 ratio and the complexes collected on a

nitrocellulose filter that allowed the free RNA (non-binders) to pass through. RNA bound to

the protein was eluted from nitrocellulose membranes by 7M urea was purified through ethanol

precipitation. Purified RNA was reverse transcribed and the cDNA was amplified by PCR

under conditions that promote low fidelity transcription.

For SELEX 1280 in each of the successive rounds the concentration of the protein was

reduced by 10% to increase the selection stringency. Nine rounds of selection were performed

of which the first 7 rounds were done with soluble trimers followed by three rounds of negative

selections and the final selection was done using GP1,2 pseudotyped VSV virus particles. The

2nd, 5th, 8th and 9th rounds were barcoded by PCR amplification using primers containing the

barcode sequences and obtained 97 million sequences from the sequencing.

Similarly, for SELEX 1281 for each successive round the concentration of protein was

reduced by 10%, 4 rounds of positive selections were performed including negative selections

against the nitrocellulose membrane. 8th and 9th round involved selection against VSV virus

particles pseudotyped with GP1,2 full length glycoprotein. Final 10th round selection was

against virus particles lacking GP1,2 to eliminate nonspecific binders. 1st, 4th, 5th, 8th and 10th

rounds were barcoded by PCR amplification and we obtained 24 million sequences from

sequencing.

Page 110: Selection of functional RNA aptamers against Ebola

102

Filter Capture Assay

To determine the binding affinity of GP1,2 recombinant proteins to RNA aptamers, 5’

32P end labeled RNA was prepared. The 2’FY modified RNA was transcribed using a

Durascribe transcription kit (Epicentre Technologies), separated by electrophoresis through an

8% polyacrylamide gel in the presence of 7 M urea, and eluted from the gel. The RNA was

dephosphorylated by incubation with calf intestine phosphatase for 1 h at 37°C, extracted with

phenol, and precipitated with ethanol. The RNA was 5’ end-labeled with [γ-32P] ATP and T4

polynucleotide kinase. In preparation to determine its binding affinity for GP1,2 and other

proteins, the RNA (2 nM) was incubated at 95°C for 5 min followed by refolding in binding

buffer [137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, 2 mM KH2PO4, 5 mM MgCl2 pH 7.4]

at room temperature (RT; 22-24 degree Celsius) for 30 min. The folded RNA was incubated

with varying concentrations of GP1,2 (10 nM-1 μM) at RT for 30 min. The RNA GP1,2

complex was captured by passing the reaction mixture through nitrocellulose membranes

(HAWP 02500) and washing with filters 3 mL of binding buffer. Quantification of

RNA/protein captured on the filter was by liquid scintillation counting. To determine the Kd

the data was fit to F=Fmin+(Fmax*L^n)/(L^n+Kd^n) (32).

Electrophoretic Mobility Shift Assay (EMSA)

Binding reactions performed with 5’ end labeled RNA that is first denatured at 95°C

for 5 min followed by slow cooling in binding buffer (137 mM NaCl, 2.7 mM KCl, 10 mM

Na2HPO4, 2 mM KH2PO4, 5 mM MgCl2, pH 7.4) for 30 min. 5uM Refolded RNA was

incubated with 0.5uM of recombinant GP1,2 delta mucin or 0.5uM BSA in binding buffer for

Page 111: Selection of functional RNA aptamers against Ebola

103

30 min at room temperature. Mixtures were analyzed for complex formation by resolving them

through a 6% (37.5:1 acrylamide:bisacrylamide) )native polyacrylamide gels run in Tris

Acetate EDTA buffer (6.75 mM TrisCl, 1 mM EDTA, 3.3 mM sodium acetate) (pH 7.4) @

200V followed by quantification of the 32P on the gel using an X ray film.

Determination of Total Protein and Percent Glycoprotein on Virus Particles

Total viral protein content was measured using Bradford assay using a standard series

of BSA concentrations 1:150 dilutions of GP1,2 pseudotyped virus particles (in duplicates).

The pseudotyped virus particles were also resolved by 10% SDS PAGE along with a known

concentration of recombinant EBOV GP1, 2 (Purchased from BPS biosciences). The gels were

stained with Coomassie blue and the bands quantified by Image J. The amount of GP1,2 in the

particles was determined relative to the control the GP1,2 control and then converted to a

percent of total protein on the virus particles using the results from the Bradford (Coomassie

blue) assay for total protein.

ACKNOWLEDGEMENT

Financial support for this study was provided by NIH grant R21AI106329-01 to MNH.

Page 112: Selection of functional RNA aptamers against Ebola

104

LEGENDS TO FIGURES

Figure 1 Comparison of top 10 sequences from SELEX experiments to select aptamers

that recognize GP1,2Δmucin and full length GP1,2

A,D) Flow charts depicting the sequential selection steps for each SELEX experiment.

Downward arrows represent positive selection steps and horizontal arrows represent negative

selections. B,E) Frequency chart of SELEX 1280 (B) and SELEX 1281 (E) depicting the

frequency of singletons, enriched and unique fraction. C,F) Frequency chart of the random

region of each round of SELEX 1280 (C) and SELEX 1280 (F)

Figure 2 Selection through a sucrose cushion

The oligonucleotides listed in the abscissa were spun through a 20% sucrose cushion in the

presence of either control VSV particles or VSV particles pseudotyped with GP1,2Δmucin.

The oligonucleotides in the pellet were sequenced to determine the fractional frequency of

each oligonucleotide. The dashed at 0.1 represents the fractional frequency expected if an

oligonucleotide were not preferentially pelleted with the particle. The asterisks identify two

oligonucleotides that preferentially pelleted with the VSV particles pseudotyped with

GP1,2Δmucin

Figure 3 Filter capture assay to estimate the affinity of oligonucleotide 4789 for

GP1,2Δmucin

A) Titration of full length GP1,2 and GP1,2Δmucin recombinant proteins against 2 nM 32P-

4789 2’FY-RNA by filter capture assay to establish the Kds, which was ND (not determined)

for GP1,2 and 143nM GP1,2Δmucin, respectively B) EMSA analysis with 5uM 32P-4789

Page 113: Selection of functional RNA aptamers against Ebola

105

2’FY-RNA and 0.5 μM BSA, 0.5 μM GP1,2 delta mucin or 64 ng/ul VSV –GP1,2Δmucin

virus particles. The arrows show the positions of the mobility shifted RNA associated with

virus particles (V) and recombinant protein (P). C) Titration of GP1,2Δmucin-pseudotyped

VSV virus particles or VSV virus particles lacking GP1,2 against 10nM 32P-4789 2’FY-RNA

by filter capture.

Figure 4 Filter capture assay to determine the affinity of oligonucleotide 5185 and 5187

for GP1,2

A) Titration of full length GP1,2 against 10 nM 32P-5185 2’FY RNA measured by the filter

capture assay from which the estimated Kd was 50nM. B) Titration of full length GP1,2

pseudotyped VSV particles against 10 nM 32P-5185 2’FY RNA measured by the filter capture

assay. C) Titration of full length GP1,2 against 10 nM 32P-5187 2’FY RNA measured by the

filter capture assay from which the estimated Kd was 560nM.

Page 114: Selection of functional RNA aptamers against Ebola

106

REFERENCES

1. Dolnik, O., Kolesnikova, L., and Becker, S. (2008) Filoviruses: Interactions with the

host cell. Cell Mol Life Sci 65, 756-776

2. Aleksandrowicz, P., Wolf, K., Falzarano, D., Feldmann, H., Seebach, J., and Schnittler,

H. (2008) Viral haemorrhagic fever and vascular alterations. Hamostaseologie 28, 77-

84

3. Feldmann, H., Jones, S., Klenk, H. D., and Schnittler, H. J. (2003) Ebola virus: from

discovery to vaccine. Nat Rev Immunol 3, 677-685

4. Noda, T., Ebihara, H., Muramoto, Y., Fujii, K., Takada, A., Sagara, H., Kim, J. H.,

Kida, H., Feldmann, H., and Kawaoka, Y. (2006) Assembly and budding of Ebolavirus.

PLoS Pathog 2, e99

5. Sanchez, A., Yang, Z. Y., Xu, L., Nabel, G. J., Crews, T., and Peters, C. J. (1998)

Biochemical analysis of the secreted and virion glycoproteins of Ebola virus. J Virol

72, 6442-6447

6. Volchkov, V. E., Feldmann, H., Volchkova, V. A., and Klenk, H. D. (1998) Processing

of the Ebola virus glycoprotein by the proprotein convertase furin. Proc Natl Acad Sci

U S A 95, 5762-5767

7. Lee, J. E., and Saphire, E. O. (2009) Ebolavirus glycoprotein structure and mechanism

of entry. Future Virol 4, 621-635

8. Manicassamy, B., Wang, J., Jiang, H., and Rong, L. (2005) Comprehensive analysis of

ebola virus GP1 in viral entry. J Virol 79, 4793-4805

Page 115: Selection of functional RNA aptamers against Ebola

107

9. Jeffers, S. A., Sanders, D. A., and Sanchez, A. (2002) Covalent modifications of the

ebola virus glycoprotein. J Virol 76, 12463-12472

10. Brindley, M. A., Hughes, L., Ruiz, A., McCray, P. B., Jr., Sanchez, A., Sanders, D. A.,

and Maury, W. (2007) Ebola virus glycoprotein 1: identification of residues important

for binding and postbinding events. J Virol 81, 7702-7709

11. Ji, X., Olinger, G. G., Aris, S., Chen, Y., Gewurz, H., and Spear, G. T. (2005) Mannose-

binding lectin binds to Ebola and Marburg envelope glycoproteins, resulting in

blocking of virus interaction with DC-SIGN and complement-mediated virus

neutralization. J Gen Virol 86, 2535-2542

12. Takada, A., Fujioka, K., Tsuiji, M., Morikawa, A., Higashi, N., Ebihara, H., Kobasa,

D., Feldmann, H., Irimura, T., and Kawaoka, Y. (2004) Human macrophage C-type

lectin specific for galactose and N-acetylgalactosamine promotes filovirus entry. J

Virol 78, 2943-2947

13. Simmons, G., Reeves, J. D., Grogan, C. C., Vandenberghe, L. H., Baribaud, F.,

Whitbeck, J. C., Burke, E., Buchmeier, M. J., Soilleux, E. J., Riley, J. L., Doms, R. W.,

Bates, P., and Pohlmann, S. (2003) DC-SIGN and DC-SIGNR bind ebola glycoproteins

and enhance infection of macrophages and endothelial cells. Virology 305, 115-123

14. Brindley, M. A., Hunt, C. L., Kondratowicz, A. S., Bowman, J., Sinn, P. L., McCray,

P. B., Jr., Quinn, K., Weller, M. L., Chiorini, J. A., and Maury, W. (2011) Tyrosine

kinase receptor Axl enhances entry of Zaire ebolavirus without direct interactions with

the viral glycoprotein. Virology 415, 83-94

Page 116: Selection of functional RNA aptamers against Ebola

108

15. Kondratowicz, A. S., Lennemann, N. J., Sinn, P. L., Davey, R. A., Hunt, C. L., Moller-

Tank, S., Meyerholz, D. K., Rennert, P., Mullins, R. F., Brindley, M., Sandersfeld, L.

M., Quinn, K., Weller, M., McCray, P. B., Jr., Chiorini, J., and Maury, W. (2011) T-

cell immunoglobulin and mucin domain 1 (TIM-1) is a receptor for Zaire Ebolavirus

and Lake Victoria Marburgvirus. Proc Natl Acad Sci U S A 108, 8426-8431

16. Kobayashi, N., Karisola, P., Pena-Cruz, V., Dorfman, D. M., Jinushi, M., Umetsu, S.

E., Butte, M. J., Nagumo, H., Chernova, I., Zhu, B., Sharpe, A. H., Ito, S., Dranoff, G.,

Kaplan, G. G., Casasnovas, J. M., Umetsu, D. T., Dekruyff, R. H., and Freeman, G. J.

(2007) TIM-1 and TIM-4 glycoproteins bind phosphatidylserine and mediate uptake of

apoptotic cells. Immunity 27, 927-940

17. Meertens, L., Carnec, X., Lecoin, M. P., Ramdasi, R., Guivel-Benhassine, F., Lew, E.,

Lemke, G., Schwartz, O., and Amara, A. (2012) The TIM and TAM families of

phosphatidylserine receptors mediate dengue virus entry. Cell Host Microbe 12, 544-

557

18. Moller-Tank, S., Kondratowicz, A. S., Davey, R. A., Rennert, P. D., and Maury, W.

(2013) Role of the phosphatidylserine receptor TIM-1 in enveloped-virus entry.

Journal of virology 87, 8327-8341

19. Miller, E. H., Obernosterer, G., Raaben, M., Herbert, A. S., Deffieu, M. S., Krishnan,

A., Ndungo, E., Sandesara, R. G., Carette, J. E., Kuehne, A. I., Ruthel, G., Pfeffer, S.

R., Dye, J. M., Whelan, S. P., Brummelkamp, T. R., and Chandran, K. (2012) Ebola

virus entry requires the host-programmed recognition of an intracellular receptor.

EMBO J 31, 1947-1960

Page 117: Selection of functional RNA aptamers against Ebola

109

20. Chandran, K., Sullivan, N. J., Felbor, U., Whelan, S. P., and Cunningham, J. M. (2005)

Endosomal proteolysis of the Ebola virus glycoprotein is necessary for infection.

Science 308, 1643-1645

21. Binning, J. M., Leung, D. W., and Amarasinghe, G. K. (2012) Aptamers in virology:

recent advances and challenges. Front Microbiol 3, 29

22. James, W. (2007) Aptamers in the virologists' toolkit. J Gen Virol 88, 351-364

23. Gopinath, S. C. (2007) Antiviral aptamers. Arch Virol 152, 2137-2157

24. Khati, M., Schuman, M., Ibrahim, J., Sattentau, Q., Gordon, S., and James, W. (2003)

Neutralization of infectivity of diverse R5 clinical isolates of human immunodeficiency

virus type 1 by gp120-binding 2'F-RNA aptamers. J Virol 77, 12692-12698

25. Biroccio, A., Hamm, J., Incitti, I., De Francesco, R., and Tomei, L. (2002) Selection of

RNA aptamers that are specific and high-affinity ligands of the hepatitis C virus RNA-

dependent RNA polymerase. J Virol 76, 3688-3696

26. Kikuchi, K., Umehara, T., Fukuda, K., Kuno, A., Hasegawa, T., and Nishikawa, S.

(2005) A hepatitis C virus (HCV) internal ribosome entry site (IRES) domain III-IV-

targeted aptamer inhibits translation by binding to an apical loop of domain IIId.

Nucleic Acids Res 33, 683-692

27. Klussman, S. (ed) (2005) The Aptamer Handbook, Wiley-VCH

28. Neff, C. P., Zhou, J., Remling, L., Kuruvilla, J., Zhang, J., Li, H., Smith, D. D.,

Swiderski, P., Rossi, J. J., and Akkina, R. (2011) An aptamer-siRNA chimera

suppresses HIV-1 viral loads and protects from helper CD4(+) T cell decline in

humanized mice. Sci Transl Med 3, 66ra66

Page 118: Selection of functional RNA aptamers against Ebola

110

29. Farokhzad, O. C., Karp, J. M., and Langer, R. (2006) Nanoparticle-aptamer

bioconjugates for cancer targeting. Expert Opin Drug Deliv 3, 311-324

30. Zhai, L., Wang, T., Kang, K., Zhao, Y., Shrotriya, P., and Nilsen-Hamilton, M. (2012)

An RNA aptamer-based microcantilever sensor to detect the inflammatory marker,

mouse lipocalin-2. Anal Chem 84, 8763-8770

31. Saijo, M., Niikura, M., Ikegami, T., Kurane, I., Kurata, T., and Morikawa, S. (2006)

Laboratory diagnostic systems for Ebola and Marburg hemorrhagic fevers developed

with recombinant proteins. Clin Vaccine Immunol 13, 444-451

32. Huang, H., Suslov, N. B., Li, N.-S., Shelke, S. A., Evans, M. E., Koldobskaya, Y., Rice,

P. A., and Piccirilli, J. A. (2014) A G-quadruplex–containing RNA activates

fluorescence in a GFP-like fluorophore. Nature chemical biology 10, 686-691

Page 119: Selection of functional RNA aptamers against Ebola

111

Figure 1

Comparison of top 10 sequences from SELEX experiments to select aptamers that

recognize GP1,2Δmucin and full length GP1,2 .

Page 120: Selection of functional RNA aptamers against Ebola

112

Figure 2

Selection through a sucrose cushion

Page 121: Selection of functional RNA aptamers against Ebola

113

Figure 3

Filter capture assay to estimate the affinity of oligonucleotide 4789 for GP1,2Δmucin

Page 122: Selection of functional RNA aptamers against Ebola

114

Figure 4

Filter capture assay to determine the affinity of oligonucleotide 5185 and 5187 for GP1,2

Page 123: Selection of functional RNA aptamers against Ebola

115

CHAPTER 4

CONCLUSION AND DISCUSSION

Ebola viruses are highly pathogenic viruses that infect humans and non-human

primates with very high mortality rates. The first Ebola virus outbreak was reported in 1976

and there have been sporadic outbreaks reported over the years. However, there still aren’t any

commercially available vaccine or therapeutic drugs against Ebola viruses. During the

devastating outbreak of Ebola Viruses in 2014, many candidate vaccines were introduced and

are still in clinical trials. In addition to vaccines, several therapeutic drugs against Ebola virus

were also expeditiously put under investigation for efficacy that are still in clinical trials. Aside

from the lack of tested therapeutics for Ebola virus, another pressing concern is the inability to

detect early virus infections. The available diagnostic kits are mostly antibody based and detect

the serum antibodies against Ebola virus, which take time to appear, or the viral RNA, which

requires correct timing of blood sampling. [1, 2]. Due to limitations, such as those just

mentioned, associated with the current detection kits there is still a need for robust, cost

effective and sensitive technique for early detection of Ebola viruses. Aptamers have been

established as a potent alternative to antibodies. Several aptamers have been selected by others

to function either as an antiviral or for biomolecule detection [3-5].

This thesis describes efforts to select and characterize 2’FY-RNA aptamers to Ebola

Glycoproteins (sGP and GP1,2) with the expectation that the aptamers will have applications

for intervention of viral host interaction and/or integration onto a biosensor platform. We chose

the full-length GP1,2 recombinant protein, GP1,2 mucin domain deleted glycoprotein and sGP

Page 124: Selection of functional RNA aptamers against Ebola

116

as the target proteins for selection to obtain aptamers that could target both the surface exposed

GP1,2 and the intracellular mucin domain deleted GP1,2. Three SELEX experiments were

performed 1280 (GP1,2Δmucin), 1281 (full-length GP1,2) and 1282 (sGP). Each SELEX

experiment was performed using an oligonucleotide pool with a 53-nucleotide random region.

Three separate pools were used and at different time intervals. Enrichment of unique clusters

was evident in all three SELEX experiments when the frequency plots of initial round pool and

final round pool was compared. A sequence comparison of the results of the three SELEX

experiments revealed identical sequences in all three. Most of the sequences in SELEX 1281

(Chapter 3) and SELEX 1282 (Chapter 2) were identical. Oligonucleotide (4797) was found

in all three selections but bound none of the targets. Thus, oligonucleotide 4797 is likely to

bind the background matrix that was common to all three SELEX experiments. One reason for

the appearance of identical sequences in 1281 and 1282 was the similar tertiary structures of

full length GP1,2 and sGP.

The observation of the same sequences appearing in two SELEX experiments was

unexpected and led to the question of the probability that same sequence could appear in two

separate DNA pools in the absence of evolution of the sequences driven by selection. The

probability of having oligonucleotides with identical sequences in separate pools is very low.

The frequency plots of the RNA pools were 99% unique, which eliminates the hypothesis that

sequences of oligonucleotides in the starting pools were redundant due to bias in the synthesis

step. Another source of bias could be in the sequencing step. The complexity of the pool is

1015 oligonucleotides representing a subset of the possible 1031 sequences and the NGS results

average a million reads per pool. Other points of bias that could increase the representation of

Page 125: Selection of functional RNA aptamers against Ebola

117

certain sequences are the transcription and PCR steps. However, none of these sources of bias

provide a credible explanation for the fact that many of the same sequences were isolated in

two independently run SELEX experiments, which started with different RNA pools. The

results could be explained by the combination of selection against target and the variations in

sequence imposed by low fidelity amplification protocol. If this is the explanation, it also

suggests that the initial selections (round 1 and 2) are very efficient in capturing

oligonucleotides with some affinity for the target. The remaining rounds of SELEX may be

dominated by the evolution of sequences due to errors in replication and transcription, which

expands the oligonucleotide populations that bind with high affinity. These results also suggest

that the selected aptamers are either unique or amongst very few possible oligonucleotides that

are capable of binding sGP with high affinity.

Chapter 2 discusses the selection of high affinity 2’F RNA aptamers that bind to sGP

with high affinity and specificity. MEME software was used to search for sequence structure

motifs in the oligonucleotides for which sequences were found in the SELEX 1281 and

SELEX1282 top clusters. We identified polyU rich sequences in the selected oligonucleotides

from both SELEX experiments. Poly U rich sequences are known to bind intracellular

regulatory RNA binding proteins [6]. We found that the oligonucleotides 5177, 5182, 5183

and 5179 with poly U rich sequences bound sGP with high affinity, whereas oligonucleotides

4789 and 5181 lacking poly U rich sequences bound with low affinity. We tested the

dependency on 2’FY modification for high affinity binding and found that unmodified 2’OH,

5183 RNA did not bind sGP. From this result, we concluded that binding of oligonucleotide

5183 to sGP depends on the 2’fluoro modification of the pyrimidines. This observation is

Page 126: Selection of functional RNA aptamers against Ebola

118

consistent with earlier reports of enhanced affinity of RNA aptamers to target proteins due to

2’F modification [7].

To identify the aptamer sequence in the larger 5183 oligonucleotide, oligonucleotides

with sequences that were truncations of the parent 5183 sequence were tested for binding to

sGP. These results confirmed the central role of the polyU-rich loop in binding sGP as the

truncated 39mer containing the hairpin U loop sequence bound sGP most effectively. This

39mer was demonstrated to bind sGP with high affinity and was given the aptamer name

39SGP1A. Our long-term goal is to integrate the selected aptamer on a detection platform.

Therefore, we tested the affinities of this aptamer for the serum proteins, human serum

albumin, fibrinogen and α2-macroglobulin.

It is important for the application of 39SGP1A to diagnostics that it not bind to other

proteins in the blood. The parent proto-aptamer, 5183 binds sGP with a Kd of 400 nM

compared with a Kd of 40 nM for sGP. The concentration of human serum albumin (HSA) in

blood is 300 μM, which means that the 5183 oligonucleotide would be saturated with HSA if

incubated with a sample of serum. However, once the sGP binding element was isolated to the

poly U stem and 5183 truncated to create the aptamer 39SGP1A there was no remaining

affinity for HSA.

In chapter 3, for selections against GP1,2 in SELEX 1280 and 1281 no conserved

sequence structural motif was found. Therefore, the affinity of oligonucleotides with sequences

in the SELEX 1280 selected families and unique sequences from SELEX 1281 were tested for

binding to GP1,2 with and without the mucin domain attached. From these tests,

oligonucleotide 4789 was found to bind GP1,2Δmucin with high affinity whereas it did not

Page 127: Selection of functional RNA aptamers against Ebola

119

bind the full length GP1,2 glycoprotein. This oligonucleotide also bound sGP (Chapter1).

Therefore, oligonucleotide 4789 appears to bind a domain that is common between sGP and

GP1,2Δmucin, which is in the N terminal portion of the GP1,2 protein sequence. The N

terminal region of GP1,2 and sGP is the conserved receptor binding domain suggesting that

4789 binds the receptor binding domain of the glycoprotein. This aptamer might have the

capability of intercepting viral host interactions. Tests of other oligonucleotides selected in

SELEX 1281for binding to the full length GP1,2 (including the mucin domain) identified two

sequences 5185 and 5187 that bound the full length GP1,2 and virus particles pseudotyped

with the full length GP1,2.

My thesis describes the development and characterization of sGP and GP1,2 specific

2’FY-modified RNA aptamers. Additional studies are needed to determine if the aptamers

selected against GP1,2 might provide therapeutic options. First, the specificity of these 2’FY-

RNA aptamers must be determined by testing their ability to bind other proteins and to bind

GP1,2 in the presence of blood proteins. The selected aptamers to full length GP1,2 could

target extracellular virus and the aptamers that recognize GP1,2 lacking the mucin domain

might target the cleaved GP1,2 protein in the endosome. Because Ebola viruses have GP1,2

independent mechanisms for entry, full length GP1,2 and GP1,2Δmucin aptamer chimeras

could be effective. The GP1,2 RNA aptamer when bound with full length GP1,2 glycoprotein

can facilitate the entry of both aptamers into the endosomes where the second aptamer to

GP1,2Δmucin could thwart viral entry into the cytoplasm. The challenge at this step could be

if these aptamers are unable to survive low pH endosomal conditions, which has not yet been

tested. An alternative approach could be to conjugate these aptamers to known therapeutic

Page 128: Selection of functional RNA aptamers against Ebola

120

drugs for intervention of viral entry. An aptamer that neutralizes EBOV will provide a means

for rapid, albeit transient, protection as a stop-gap measure prior to further treatment including

isolation of the infected individual from others who might be susceptible.

Our future goal for the selected aptamer 39SGP1A is to use it to detect sGP. The

aptamer will be integrated onto microcantilever a platform that is equipped with a system to

detect RNA aptamer-sGP interaction with high sensitivity and specificity such as was

previously demonstrated with mouse lipocalin-2 [8] . Although, we have tested the aptamers

for its specificity against human serum albumin, there are many known proteins present at

various concentrations in the serum. Therefore, binding of these aptamers to sGP in the

presence of serum should be evaluated. Aptamer efficiency is also determined by its tertiary

folding, therefore following the integration of the selected aptamer onto a platform, it must be

ensured that folding of the aptamer hasn’t been compromised.

In summary, the work presented in this thesis focuses on the selection and

characterization of 2’FY-RNA aptamers against Ebola viral glycoproteins. Results from the

selections reveal various important factors responsible for the selection of high affinity

aptamers. These include considerations with respect to designing the pool, ensuring evolution

of oligonucleotide sequences and the importance of the 2’F modification in mediating high

affinity interactions with proteins. These studies resulted in the identification of aptamers that

bind sGP and GP1,2, which can be developed as antivirals or for detection of Ebola virus.

Page 129: Selection of functional RNA aptamers against Ebola

121

REFERENCES

1. Ksiazek, T.G., et al., ELISA for the detection of antibodies to Ebola viruses. The Journal

of infectious diseases, 1999. 179(Supplement_1): p. S192-S198.

2. Su, S., et al., Diagnostic strategies for Ebola virus detection. The Lancet Infectious

Diseases, 2016. 16(3): p. 294-295.

3. Gopinath, S.C., Antiviral aptamers. Arch Virol, 2007. 152(12): p. 2137-57.

4. Bruno, J.G., Predicting the Uncertain Future of Aptamer-Based Diagnostics and

Therapeutics. Molecules, 2015. 20(4): p. 6866-6887.

5. Szpechciński, A. and A. Grzanka, [Aptamers in clinical diagnostics]. Postepy

biochemii, 2005. 52(3): p. 260-270.

6. Kanai, A., K. Tanabe, and M. Kohara, Poly (U) binding activity of hepatitis C virus

NS3 protein, a putative RNA helicase. FEBS letters, 1995. 376(3): p. 221-224.

7. DEY, A.K., et al., Structural characterization of an anti-gp120 RNA aptamer that

neutralizes R5 strains of HIV-1. Rna, 2005. 11(6): p. 873-884.

8. Zhai, L., et al., An RNA aptamer-based microcantilever sensor to detect the

inflammatory marker, mouse lipocalin-2. Anal Chem, 2012. 84(20): p. 8763-70.