bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

1 Spike mutation T403R allows bat

2 RaTG13 to use human ACE2

3 Fabian Zech1, Daniel Schniertshauer1, Christoph Jung2,3,4, Alexandra Herrmann5, Qinya 4 Xie1, Rayhane Nchioua1, Caterina Prelli Bozzo1, Meta Volcic1, Lennart Koepke1, Jana 5 Krüger6, Sandra Heller6, Alexander Kleger6, Timo Jacob2,3,4, Karl-Klaus Conzelmann7, 6 Armin Ensser5, Konstantin M.J. Sparrer1 and Frank Kirchhoff1*

7

8 1Institute of Molecular Virology, Ulm University Medical Center, 89081 Ulm, Germany; 9 2Institute of Electrochemistry, Ulm University, 89081 Ulm, Germany; 3Helmholtz-Institute 10 Ulm (HIU) Electrochemical Energy Storage, Helmholtz-Straße 16, 89081 Ulm, Germany; 11 4Karlsruhe Institute of Technology (KIT), P.O. Box 3640, 76021 Karlsruhe, Germany; 12 5Institute of Clinical and Molecular Virology, University Hospital Erlangen, Friedrich- 13 Alexander Universität Erlangen-Nürnberg, 91054 Erlangen, Germany. 6Department of 14 Internal Medicine I, Ulm University Medical Center, 89081 Ulm, Germany; 7Max von 15 Pettenkofer-Institute of Virology, Medical Faculty, and Gene Center, Ludwig-Maximilians- 16 Universität München, 81377 Munich, Germany. 17 18 * Address Correspondence to: [email protected] 19

20 Running title: T403R allows RaTG13 Spike to use human ACE2 21 22 KEYWORDS: SARS-CoV-2, RaTG13, spike glycoprotein, ACE2 receptor, viral zoonosis

1

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

23 ABSTRACT

24 Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), the cause of the

25 COVID-19 pandemic, most likely emerged from bats1. A prerequisite for this devastating

26 zoonosis was the ability of the SARS-CoV-2 Spike (S) glycoprotein to use human

27 angiotensin-converting enzyme 2 (ACE2) for viral entry. Although the S protein of the

28 closest related bat , RaTG13, shows high similarity to the SARS-CoV-2 S protein it

29 does not efficiently interact with the human ACE2 receptor2. Here, we show that a single

30 T403R mutation allows the RaTG13 S to utilize the human ACE2 receptor for infection

31 of human cells and intestinal organoids. Conversely, mutation of R403T in the SARS-

32 CoV-2 S significantly reduced ACE2-mediated virus infection. The S protein of SARS-

33 CoV-1 that also uses human ACE2 also contains a positive residue (K) at this position,

34 while the S proteins of CoVs utilizing other receptors vary at this location. Our results

35 indicate that the presence of a positively charged amino acid at position 403 in the S

36 protein is critical for efficient utilization of human ACE2. This finding could help to

37 predict the zoonotic potential of animal .

2

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

38 Since its first occurrence in Wuhan in December 2019, SARS-CoV-2, the causative agent of

39 COVID-19, has infected about 170 million people by May 2021 and caused a global health and

40 economic crisis3. SARS-CoV-2 belongs to the Sarbecovirus subgenus of ,

41 which are mainly found in bats3. Horseshoe bats (Rhinolophidae) also harbour that are

42 closely related to SARS-CoV-1 that infected about 8.000 people in 2002 and 20033,4. The bat

43 virus RaTG13 sampled from a Rhinolophus affinis horseshoe bat in 2013 in has been

44 identified as the closest relative of SARS-CoV-2 showing approximately 96% sequence

45 identity throughout the genome1. Thus, SARS-CoV-2 most likely originated from horseshoe

46 bats1,5, although it has been proposed that cross-species transmission to humans may have

47 involved as secondary intermediate host6,7.

48 The Spike (S) proteins of both SARS-CoV-1 and SARS-CoV-2 utilize the human

49 angiotensin-converting enzyme 2 (ACE2) receptor to enter human target cells8–11. The ability

50 to use a human receptor for efficient infection is a key prerequisite for successful zoonotic

51 transmission. Although the RaTG13 S protein is highly similar to the SARS-CoV-2 S it does

52 not interact efficiently with the human ACE2 receptor2, suggesting that this bat virus would

53 most likely not be able to directly infect humans. It has been reported that specific alterations

54 in the receptor-binding domain (RBD)12, as well as a four-amino-acid insertion (PRRA)

55 introducing a furin-cleavage site,8,13 play a key role in efficient ACE2 utilization and

56 consequently the high infectiousness and efficient spread of SARS-CoV-2. However, it

57 remains poorly understood which specific features allow the S proteins of bat CoVs to use

58 human ACE2 as entry cofactor and thus to successfully cross the species barrier to humans.

59 Computational analyses suggested that R403 is involved in intramolecular interactions

60 stabilizing the SARS-CoV-2 S trimer interface2 and contributes significantly to the strength of

61 SARS-CoV-2 RBD interaction with the human ACE2 receptor14. We found that R403 is highly

62 conserved in SARS-CoV-2 S proteins: only 233 of 1.7 million sequence records contain a

63 conservative change of R403K and just 18 another residue or deletion. Notably, the presence

3

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

64 of a positively charged residue at position 403 distinguishes the S proteins of SARS-CoV-1

65 (K403) and SARS-CoV-2 (R403) from the bat CoV RaTG13 S protein (T403) (Fig. 1a).

66 Molecular modelling of S/ACE2 interaction using reactive force field simulations confirmed

67 close proximity and putative charge interactions between R403 in the SARS-CoV-2 S with E37

68 in the human ACE2 receptor (Fig. 1b). These analyses predicted that mutation of T403R

69 significantly strengthens the ability of the RaTG13 S protein to bind human ACE2 (Fig. 1c,

70 Extended Data Movies 1 and 2).

71 To verify the functional importance of residue 403 for ACE2 usage by CoV S proteins, we

72 used VSV particles (VSVpp) pseudotyped with parental and mutant S proteins. Mutation of

73 R403T reduced the ability of the SARS-CoV-2 S protein to mediate entry of VSVpp into Caco-

74 2 cells by 40% (Fig. 2a, b). Strikingly, the T403R change enhanced the infectiousness of

75 VSVpp carrying the RaTG13 S ~30-fold, while substitution of T403A introduced as control

76 had no enhancing effect (Fig. 2b). Cell-to-cell fusion assays showed that coexpression of the

77 SARS-CoV-2 S and human ACE2 resulted in the formation of large syncytia (Extended Data

78 Fig. 1). The parental and T403A RaTG13 S did not lead to significant fusion but significant

79 syncytia formation was observed for the T403R RaTG13 S (Extended Data Fig. 1). Western

80 blot analyses showed that the mutant S proteins were efficiently expressed and incorporated

81 into VSVpp, albeit the SARS-CoV-2 R403T S with reduced efficiency (Extended Data Fig.

82 2). In line with the VSVpp results, complementation of a full-length recombinant SARS-CoV-

83 2 lacking the S ORF (SCoV-2ΔS) in ACE2-expressing HEK293T cells with wildtype (WT)

84 SARS-CoV-2 S led to virus-induced cytopathic effects (CPE) indicating successful virus

85 production and propagation (Fig. 2c). Mutation of R403T in the SARS-CoV-2 S reduced CPE.

86 The WT and T403A RaTG13 S were entirely unable to complement SCoV-2ΔS, while the

87 T403R RaTG13 S resulted in significant CPE. Expression of a Gaussia luciferase (GLuc) from

88 S variant complemented recombinant SCoV2ΔS-GLuc confirmed the importance of R403 for

89 viral spread (Fig. 2d).

4

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

90 Coronavirus entry is a multi-step process and critically dependent on proteolytic processing

91 of the S protein15. The interaction of the SARS-CoV-2 S trimer with ACE2 promotes

92 proteolytic processing16,17. Western blot analysis revealed that ACE2 coexpression induces

93 efficient cleavage of the SARS-CoV-2 and T403R RaTG13 S proteins to S2, while cleavage

94 of the WT and T403A RaTG13 S proteins remained inefficient (Extended Data Fig. 3). R403

95 generates a potential RGD integrin binding site in the viral Spike protein and it is under debate

96 whether the ability of the SARS-CoV-2 S to use integrins as viral attachment factors may play

97 a role in its high infectiousness18,19. The integrin inhibitor ATN-161 had no significant effect

98 on SARS-CoV-2 or T403R RaTG13 S-mediated infection (Extended Data Fig. 4a, b). Thus,

99 the enhancing effect of the T403R mutation on the ability of RaTG13 S to infect human cells

100 seems to be due to increased interaction with ACE2 rather than utilization of integrins. Taken

101 together, our results demonstrate that mutation of T403R strongly enhances the ability of the

102 bat RaTG13 S protein to utilize ACE2 for infection of human cells.

103 To assess whether the T403R change might allow the bat CoV RaTG13 to spread to different

104 human organs, we performed infection studies using intestinal organoids derived from

105 pluripotent stem cells. The parental SARS-CoV-2 S protein allowed efficient infection of gut

106 organoids20 and the R403T change had modest attenuating effects (Fig. 3, Extended Data Fig.

107 5). In contrast, the parental RaTG13 S protein did not result in significant VSVpp infection,

108 while the corresponding T403R mutant allowed significant infection of human intestinal cells

109 (Fig. 3; Extended Data Fig. 5).

110 To examine the species-specificity of receptor usage by SARS-CoV-2 and RaTG13 S

111 proteins, we overexpressed human and bat derived ACE2 in HEK293T cells and examined

112 their susceptibility to S-mediated VSVpp infection. The WT SARS-CoV-2 and the T403R

113 RaTG13 S proteins allowed efficient entry into cells overexpressing human ACE2, while the

114 parental RaTG13 S protein was poorly active (Fig. 4a). Both WT SARS-CoV-2 S and (to a

115 lesser extent) R403T SARS-CoV-2 S proteins were also capable of using bat (Rhinolophus

5

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

116 affinis) ACE2 for viral entry although the overall infection rates were low (Fig. 4a, Extended

117 Data Fig. 6). In contrast, the RaTG13 S proteins were unable to use bat ACE2 for infection

118 suggesting that RaTG13 might use an alternative receptor for infection of bat cells. The results

119 agree with the previous finding that RaTG13 S is able to use human ACE2 to some extent if

120 overexpressed21 but further demonstrate that the T403R greatly enhances this function and is

121 required for utilization of endogenously expressed human ACE2.

122 To validate the results obtained with human HEK293T cells, we utilized the lung epithelial

123 cell line Tb1 Lu1 of Tadarida brasiliensis (Bat31)22. In agreement with the previous finding

124 that this cell line lacks endogenous ACE2 expression, it did not support infection by CoV S

125 proteins (Fig. 4b). Engineered expression of human ACE2 rendered Lu 1 highly susceptible to

126 infection mediated by SARS-CoV-2 and the T403R RaTG13 S proteins (Fig. 4b). In

127 comparison, entry via the R403T SARS-CoV-2 S was strongly attenuated and the parental and

128 T403A RaTG13 S proteins were unable to mediate significant VSV-pp infection.

129 Our results demonstrate that a single amino acid change of T403R allows RaTG13, the

130 closest known bat relative of SARS-CoV-2, to utilize human ACE2 for viral entry. The strong

131 enhancing effect of the T403R change on RaTG13 S function came as surprise since five of six

132 different residues proposed to be critical for SARS-CoV-2 S RBD interaction with human

133 ACE2 are not conserved in RaTG13 S12,23. A very recent study proposed that residue 501 plays

134 a key role in the ability of RaTG13 S to use human ACE2 for viral entry24 but the reported

135 enhancing effect of changes at position 501 was weaker than that observed for the T403R

136 change analysed in the present study. However, the previous finding that numerous residues in

137 the SARS-CoV-2 S RBD are involved in the interaction with the human ACE2 orthologue

138 explains why the R403T substitution only moderately reduced SARS-CoV-2 infection. It has

139 been shown that the RBD of SARS-CoV-2 S shows higher homology to the corresponding

140 region of the CoV S protein than to RaTG136,7. Whether or not this is a consequence

141 of recombination or convergent evolution is under debate25,26. Notably, the Pan CoV-S protein

6

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

142 also contains a positive residue (K) at position 403 (Fig. 1a) and is capable of utilizing human

143 ACE2 for infection. Altogether our results suggest that a positive residue at position 403 in the

144 S protein was most likely a prerequisite for efficient zoonotic transmission and pandemic

145 spread of SARS-CoV-2. We found that a positively charged residue at the corresponding

146 position is present in the S proteins of the great majority of RaTG13-related bat coronaviruses

147 (Extended Data Fig. 7) raising the possibility that many bat sarbecoviruses, including the

148 unknown precursor of SARC-CoV-2, are fitter for zoonotic transmission than RaTG13.

149

150 Methods

151 Molecular dynamics simulation. Based on the structure of ACE2-bounded to SARS-CoV-2

152 taken from the Protein Data Bank27 (identification code 7KNB), the initial atomic positions

153 were obtained. Equilibration (300K for 0.5 ns) was performed by ReaxFF (reactive molecular

154 dynamic) simulations28 within the Amsterdam Modeling Suite 2020 (ADF2020, SCM,

155 Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands,

156 http://www.scm.com). Based on the equilibrated structure, amino acids from the spike protein

157 were replaced with the respective amino acids from RaTG13, respectively the modification.

158 After an additional equilibration (300K for 0.5 ns) ReaxFF (reactive molecular dynamic)

159 simulations were performed within the NVT ensemble over 25 ps, while coupling the system

160 to a Berendsen heat bath (T=300 K with a coupling constant of 100 fs). The interaction energy

161 was obtained by averaging over these simulations. For all visualizations the Visual Molecular

162 Dynamics program (VMD)29 was used.

163 Cell culture and viruses. All cells were cultured at 37°C in a 5% CO2 atmosphere. Human

164 embryonic kidney 293T cells purchased from American type culture collection (ATCC:

165 #CRL3216) were cultivated in Dulbecco’s Modified Eagle Medium (DMEM, Gibco)

166 supplemented with 10% (v/v) heat-inactivated fetal bovine serum (FBS, Gibco), 2 mM L-

7

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

167 glutamine (PANBiotech), 100 µg/ml streptomycin (PANBiotech) and 100 U/ml penicillin

168 (PANBiotech). Calu-3 (human epithelial lung adenocarcinoma, kindly provided and verified

169 by Prof. Frick, Ulm University) cells were cultured in Minimum Essential Medium Eagle

170 (MEM, Sigma) supplemented with 10% (v/v) FBS (Gibco) (during viral infection) or 20%

171 (v/v) FBS (Gibco) (during all other times), 100 U/ml penicillin (PAN-Biotech), 100 µg/ml

172 streptomycin (PAN-Biotech), 1 mM sodium pyruvate (Gibco), and 1 mM NEAA (Gibco).

173 Caco-2 (human epithelial colorectal adenocarcinoma, kindly provided by Prof. Holger Barth,

174 Ulm University) cells were cultivated in DMEM (Gibco) containing 10% FBS (Gibco), 2 mM

175 glutamine (PANBiotech), 100 µg/ml streptomycin (PANBiotech), 100 U/ml penicillin

176 (ANBiotech), 1 mM Non-essential amino acids (NEAA, Gibco), 1 mM sodium pyruvate

177 (Gibco). I1-Hybridoma cells were purchased from ATCC (#CRL-2700) and cultured in RPMI

178 supplemented with 10% (v/v) heat-inactivated FBS (Gibco), 2 mM L-glutamine

179 (PANBiotech), 100 µg/ml streptomycin (PANBiotech) and 100 U/ml penicillin (PANBiotech).

180 Tb 1 Lu (Tadarida brasiliensis derived lung epithelial) and Ri 1 Lu huACE2 (Rhinolophus

181 affinis derived lung epithelial cells expressing human ACE2, ACE2, kindly provided by Marcel

182 A. Müller, were cultured in DMEM supplemented with 10% (v/v) heat-inactivated FBS

183 (Gibco), 2 mM L-glutamine (PANBiotech), 100 µg/ml streptomycin (PANBiotech) and 100

184 U/ml penicillin (PANBiotech), 2 mM sodium pyruvate (Gibco). Viral isolate

185 BetaCoV/France/IDF0372/2020 (#014V-03890) was obtained through the European Virus

186 Archive global.

187 Expression constructs. pCG_SARS-CoV-2-Spike-IRES_eGFP, coding the spike protein of

188 SARS-CoV-2 isolate Wuhan-Hu-1, NCBI reference Sequence YP_009724390.1, was kindly

189 provided by Stefan Pöhlmann (German Primate Center, 473 Göttingen, Germany).

190 pCG_SARS-CoV-2-Spike C-V5-IRES_eGFP and RaTG13-S (synthesized by Baseclear) was

191 PCR amplified and subcloned into a pCG-IRES_eGFP expression construct using the

192 restriction enzymes XbaI and MluI (New England Biolabs). The SARS-CoV-2 S R403T and

8

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

193 RaTG13 S T403R/T403A mutant plasmids were generated using Q5 Site-Directed

194 Mutagenesis Kit (NEB).

195 Cloning of SARS-CoV-2 ΔS bacmid. An anonymized residual respiratory swab sample from

196 a patient with SARS-CoV-2 infection was used as a template for genome amplification. Total

197 nucleic acids were extracted on an automated Qiagen EZ1 robotic workstation using the Qiagen

198 EZ1 virus mini kit v2.0 according to the manufacturer's instructions. Genomic viral RNA was

199 reverse transcribed using the NEB LunaScript RT SuperMix Kit according to the

200 manufacturer’s protocol. Four overlapping fragments covering the entire viral genome were

201 amplified using the NEB Q5 High-Fidelity DNA Polymerase. The resulting amplicons were

202 assembled with a modified pBeloBAC11 backbone, containing CMV and T7 promotors as well

203 as the HDV ribozyme and bGH polyA signal, using the NEBuilder HiFi DNA Assembly

204 Cloning Kit. Assembled DNA was electroporated into E. coli GS1783 strain and resulting

205 clones of pBelo-SARSARS-CoV-2 were confirmed by restriction digestion and next

206 generation sequencing. The viral Spike gene was replaced with a kanamycin-cassette flanked

207 by SacII restriction sites by homologous recombination using the Lambda-Red Recombination

208 System30. The bacmid was linearized with the restriction enzyme SacII, and EGFP or GLuc

209 reporter cassettes were introduced instead of Spike using the the NEBuilder HiFi DNA

210 Assembly Cloning Kit according to the manufacturer’s instruction. Positive clones were

211 confirmed by restriction digestion and sequencing.

212 SARS-CoV-2 ΔS replicon system. HEK293 T cells were seeded in six well format and

213 transfected with 3 µg pBelo-SARSCoV-2-dSpike-GLuc-K2 or pBelo-SARSCoV-2-dSpike-

214 EGFP and 0.25 µg of each expression construct pLVX-EF1alpha-SARS-CoV2-N-2xStrep-

215 IRES-Puro, pCG-ACE2, pCAG-T7-RNA-polymerase and one pCG- vector encoding the spike

216 protein of SARS-CoV-2, RaTG13 or the indicated mutant S respectively. Two days after

217 transfection, bright field and fluorescence microscopy (GFP) images were acquired using the

218 Cytation 3 microplate reader (BioTek). Gaussia luciferase activity in the supernatants was

9

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

219 measured with the Gaussia Luciverase Assay system (Promega) according to the company’s

220 instructions.

221 Transfections. Plasmid DNA was transfected using either calcium phosphate transfection or

222 Polyethylenimine (PEI, 1 mg/ml in H2O, Sigma-Aldrich) according to the manufacturers

223 recommendations or as described previously31.

224 Pseudoparticle production. To produce pseudotyped VSVΔG-GFP particles, 6*106 HEK 293

225 T cells were seeded 18 hours before transfection in 10 cm dishes. The cells were transfected

226 with 15 µg of a glycoprotein expressing vector using PEI (PEI, 1 mg/ml in H2O, Sigma-

227 Aldrich). Twenty-four hours post transfection, the cells were infected with VSVΔG-GFP

228 particles pseudotyped with VSV G at a MOI of 3. One hour post-infection, the inoculum was

229 removed. Pseudotyped VSVΔG-GFP particles were harvested 16 hours post infection. Cell

230 debris were pelleted and removed by centrifugation (500 g, 4 °C, 5 min). Residual input

231 particles carrying VSV-G were blocked by adding 10 % (v/v) of I1 Hybridoma Supernatant

232 (I1, mouse hybridoma supernatant from CRL-2700; ATCC) to the cell culture supernatant.

233 Whole-cell and cell free lysates. Whole-cell lysates were prepared by collecting cells in

234 Phosphate-Buffered Saline (PBS, Gibco), pelleting (500 g, 4 °C, 5 min), lysing and clearing as

235 previously described31. Total protein concentration of the cleared lysates was measured using

236 the Pierce BCA Protein Assay Kit (Thermo Scientific) according to manufacturer’s

237 instructions. Viral particles were filtered though a 0.45 µm MF-Millipore Filter (Millex) and

238 centrifuged through a 20% sucrose (Sigma) cushion. The pellet was lysed in transmembrane

239 lysis buffer already substituted with Protein Sample Loading Buffer (LI-COR).

240 SDS-PAGE and immunoblotting. SDS-PAGE and immunoblotting was performed as

241 previously described31. In brief, whole cell lysates were mixed with 4x Protein Sample Loading

242 Buffer (LI-COR, at a final dilution of 1x) supplemented with 10% (v/v) Tris(2-

243 Carboxyethyl)phosphine hydrochloride 0.5 M (SIGMA), heated to 95°C for 10 min separated

10

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

244 on NuPAGE 4-12% Bis-Tris Gels (Invitrogen) for 90 min at 120 V and blotted at constant 30

245 V for 30 min onto Immobilon-FL PVDF membrane (Merck Millipore). After the transfer, the

246 membrane was blocked in 1% Casein in PBS (Thermo Scientific) and stained using primary

247 antibodies directed against SARS-CoV-2 S (1:1,000, Biozol, 1A9, #GTX632604), ACE2

248 (1:1,000, Abcam, #GTX632604), VSV-M (1:2,000, Absolute Antibody, 23H12, #Ab01404-

249 2.0), V5-tag (1:1,000, Cell Signaling, #13202), GAPDH (1:1,000, BioLegend, #631401) and

250 Infrared Dye labelled secondary antibodies (1:20,000, LI-CORIRDye). Proteins were detected

251 using a LI-COR Odyssey scanner and band intensities were quantified using LI-COR Image

252 Studio.

253 Stem Cell Culture and Intestinal Differentiation. Human embryonic stem cell line HUES8

254 (Harvard University, Cambridge, MA) was used with permission from the Robert Koch

255 Institute according to the “79. Genehmigung nach dem Stammzellgesetz, AZ 3.04.02/0084.”

256 Cells were cultured on human embryonic stem cell matrigel (Corning, Corning, NY) in mTeSR

257 Plus medium (STEMCELL Technologies, Vancouver, Canada) at 5% CO2, 5% O2, and 37°C.

258 Medium was changed every other day and cells were split with TrypLE Express (Invitrogen,

259 Carlsbad, CA) twice a week. For differentiation, 300,000 cells per well were seeded in 24-well

260 plates coated with growth factor–reduced matrigel (Corning) in mTeSR Plus with 10 mM Y-

261 27632 (STEMCELL Technologies). The next day, differentiation was started at 80%-90%

262 confluency, as described previously32.

263 Intestinal organoids. To prepare in vitro differentiated organoids for transduction, matrigel

264 was dissolved in Collagenase/Dispase (Roche, Basel, Switzerland) for 2 hours at 37°C and

265 stopped by cold neutralization solution (DMEM, 1% bovine serum albumin, and 1% penicillin-

266 streptomycin). Organoids were transferred into 1.5-mL tubes and infected in 300 μL

267 pseudoparticle containing inoculum. Organoids were then resuspended in 35-μL cold growth

268 factor–reduced matrigel to generate cell-matrigel domes in 48-well plates. After 10 minutes at

269 37°C, intestinal growth medium (DMEM F12 [Gibco, Gaithersburg, MD], 1× B27 supplement 11

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

270 [Thermo Fisher Scientific], 2-mM L-glutamine, 1% penicillin-streptomycin, 40 mM HEPES

271 [Sigma-Aldrich], 3 μM CHIR99021, 200 nM LDN-193189 [Sigma-Aldrich], 100 ng/mL hEGF

272 [Novoprotein, Summit, NJ], and 10 μM Y-27632 [STEMCELL Technologies]) was added and

273 organoids were incubated at 37°C. The Organoids were imaged using the Cytation 3 cell

274 imaging system and processed with Gen 5 and ImageJ software. For FACS preparation, the

275 matrigel was dissolved and the extracted organoids were dissolved in Accutase (Stemcell

276 technologies). The cells were fixed with PBS for 10 min at 4°C and washed with cold PBS

277 containing 2% FBS. Flow cytometry analyses were performed using a FACS CANTO II (BD)

278 flow cytometer. Transduction rates were determined by GFP expression and analysed with

279 DIVA and Flowjow10 software.

280 α5β5 integrin blocking. Caco-2 cells were preincubated with the indicated amounts of α5β5

281 integrin Inhibitor ATN-161 (Sigma) for two hours and infected with 100 µl freshly produced

282 VSVΔG-GFP pseudo particles. 16 hours post infection, GFP positive cells were automatically

283 quantification using a Cytation 3 microplate reader (BioTek). Calu-3 cells were preincubated

284 with the indicated amounts of ATN-161 (Sigma) for two hours and infected with SARS-CoV-

285 2 Viral isolate BetaCoV/France/IDF0372/2020 (MOI 0.05, six hours). 48 hours post-infection

286 supernatants were harvested for qRT-PCR analysis.

287 Sequence Logo and alignments. Alignments of primary bat sequences (GQ153541.1/1-

288 71,GQ153544.1/1-71,GQ153540.1, GQ153539.1, DQ084200.1, DQ084199.1, GQ153548.1,

289 GQ153547.1, GQ153546.1, GQ153545.1, DQ022305.2, GQ153542.1, GQ153543.1,

290 KJ473815.1, KF294457.1, KY417148.1, KJ473814.1, MK211374.1, KY417142.1,

291 MK211377.1, JX993988.1, DQ412043.1, DQ648857.1, JX993987.1, KY417143.1,

292 KY417147.1, MK211378.1, DQ648856.1, KJ473812.1, KY770860.1, KY770858.1,

293 KY770859.1, KJ473816.1, RmYN02, KY417145.1, KU182964.1, KY938558.1, KJ473811.1,

294 KJ473813.1, MG772933.1, MG772934.1, KY417150.1, KT444582.1, KY417152.1,

295 MK211376.1, GU190215.1, MN996532.1, EF065513.1, MG693170.1, MG762674.1,

12

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

296 HM211101.1, HM211099.1, EF065514.1, EF065516.1, EF065515.1, MK492263.1,

297 MG693168.1, MG693172.1, MG693169.1, MG693171.1, KU762337.1, KU762338.1,

298 HQ166910.1, KT253270.1, KT253269.1, KY073748.1, MN611517.1, KY073747.1,

299 KY073744.1, KY073745.1, KY073746.1, NC_028833.1, MK720944.1, NC_010437.1/1-

300 7,EU420138.1, KJ473796.1, MN611524.1, KJ473795.1, EU420137.1, KJ473799.1,

301 KJ473800.1, KJ473797.1, MN611518.1, KY770850.1, KY770851.1, KJ473806.1,

302 EU420139.1, KJ473798.1, MG916902.1, MG916903.1, JQ989269.1, JQ989267.1,

303 JQ989268.1, JQ989266.1, JQ989272.1, JQ989273.1, MN611523.1, MN611525.1,

304 JQ989271.1, JQ989270.1, MK720945.1, MK720946.1, MG916904.1, KJ473810.1,

305 NC_028814.1, DQ648858.1, NC_009657.1, MN611521.1, KF430219.1, NC_009988.1/1-

306 7,EF203066.1, EF203067.1, EF203065.1, MF370205.1, KJ473808.1, MN611522.1,

307 DQ648794.1, EF065505.1, EF065506.1, EF065508.1, MH002339.1, MN611519.1,

308 MH002338.1, KJ473822.1, MH002337.1, KU182965.1, EF065507.1, EF065510.1,

309 EF065511.1, EF065512.1, MH002342.1, EF065509.1, KJ473820.1, MH002341.1,

310 MN611520.1, KX442565.1, KX442564.1/1-71) was performed using ClustalW33 with a

311 gapOpening penalty of 80. Sequence logos were generated using R packages ggplot2 and

312 ggseqlogo34.

313 Statistics. Statistical analyses were performed using GraphPad PRISM 8 (GraphPad Software).

314 P-values were determined using a two-tailed Student’s t test with Welch’s correction. Unless

315 otherwise stated, data are shown as the mean of at least three independent experiments ± SEM.

316 References

317 1. Zhou, P. et al. A outbreak associated with a new coronavirus of probable 318 bat origin. Nature 579, 270–273 (2020). 319 2. Wrobel, A. G. et al. SARS-CoV-2 and bat RaTG13 spike glycoprotein structures 320 inform on virus evolution and furin-cleavage effects. Nat Struct Mol Biol 27, 763–767 321 (2020). 322 3. Li, W. et al. Bats Are Natural Reservoirs of SARS-Like Coronaviruses. Science 310, 323 676–679 (2005). 13

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

324 4. Hu, B. et al. Discovery of a rich gene pool of bat SARS-related coronaviruses 325 provides new insights into the origin of SARS coronavirus. PLoS Pathog 13, e1006698 326 (2017). 327 5. Wu, F. et al. A new coronavirus associated with human respiratory disease in . 328 Nature 579, 265–269 (2020). 329 6. Lam, T. T. Y. et al. Identifying SARS-CoV-2 related coronaviruses in Malayan 330 pangolins. Nature (2020) doi:10.1038/s41586-020-2169-0. 331 7. Xiao, K. et al. Isolation of SARS-CoV-2-related coronavirus from Malayan 332 pangolins. Nature (2020) doi:10.1038/s41586-020-2313-x. 333 8. Walls, A. C. et al. Structure, Function, and Antigenicity of the SARS-CoV-2 Spike 334 Glycoprotein. Cell 181, 281-292.e6 (2020). 335 9. Letko, M., Marzi, A. & Munster, V. Functional assessment of cell entry and receptor 336 usage for SARS-CoV-2 and other lineage B betacoronaviruses. Nature Microbiology 5, 562– 337 569 (2020). 338 10. Hoffmann, M. et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and 339 Is Blocked by a Clinically Proven Protease Inhibitor. Cell (2020) 340 doi:10.1016/j.cell.2020.02.052. 341 11. Li, W. et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS 342 coronavirus. Nature 426, 450–454 (2003). 343 12. Wan, Y., Shang, J., Graham, R., Baric, R. S. & Li, F. Receptor Recognition by the 344 Novel Coronavirus from Wuhan: an Analysis Based on Decade-Long Structural Studies of 345 SARS Coronavirus. J Virol 94, (2020). 346 13. Hoffmann, M., Kleine-Weber, H. & Pöhlmann, S. A Multibasic Cleavage Site in the 347 Spike Protein of SARS-CoV-2 Is Essential for Infection of Human Lung Cells. Molecular 348 Cell 78, 779–784.e5 (2020). 349 14. Laurini, E., Marson, D., Aulic, S., Fermeglia, A. & Pricl, S. Computational 350 Mutagenesis at the SARS-CoV-2 Spike Protein/Angiotensin-Converting Enzyme 2 Binding 351 Interface: Comparison with Experimental Evidence. ACS Nano 15, 6929–6948 (2021). 352 15. Li, F. Structure, Function, and Evolution of Coronavirus Spike Proteins. Annu Rev 353 Virol 3, 237–261 (2016). 354 16. Raghuvamsi, P. V. et al. SARS-CoV-2 S protein:ACE2 interaction reveals novel 355 allosteric targets. eLife 10, e63646 (2021). 356 17. Beniac, D. R., deVarennes, S. L., Andonov, A., He, R. & Booth, T. F. Conformational 357 reorganization of the SARS coronavirus spike following receptor binding: implications for 358 membrane fusion. PLoS One 2, e1082 (2007). 359 18. Sigrist, C. J., Bridge, A. & Le Mercier, P. A potential role for integrins in host cell 360 entry by SARS-CoV-2. Antiviral Res 177, 104759 (2020). 361 19. Othman, H. et al. SARS-CoV-2 spike protein unlikely to bind to integrins via the 362 Arg-Gly-Asp (RGD) motif of the Receptor Binding Domain: evidence from structural

14

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

363 analysis and microscale accelerated molecular dynamics. bioRxiv 2021.05.24.445335 (2021) 364 doi:10.1101/2021.05.24.445335. 365 20. Krüger, J. et al. Drug Inhibition of SARS-CoV-2 Replication in Human Pluripotent 366 Stem Cell-Derived Intestinal Organoids. Cell Mol Gastroenterol Hepatol (2020) 367 doi:10.1016/j.jcmgh.2020.11.003. 368 21. Shang, J. et al. Structural basis of receptor recognition by SARS-CoV-2. Nature 581, 369 221–224 (2020). 370 22. Hoffmann, M. et al. Differential Sensitivity of Bat Cells to Infection by Enveloped 371 RNA Viruses: Coronaviruses, Paramyxoviruses, Filoviruses, and Influenza Viruses. PLoS 372 One 8, (2013). 373 23. Andersen, K. G., Rambaut, A., Lipkin, W. I., Holmes, E. C. & Garry, R. F. The 374 proximal origin of SARS-CoV-2. Nat Med 26, 450–452 (2020). 375 24. Liu, K. et al. Binding and molecular basis of the bat coronavirus RaTG13 virus to 376 ACE-2 in humans and other species. Cell (2021) doi:10.1016/j.cell.2021.05.031. 377 25. Boni, M. F. et al. Evolutionary origins of the SARS-CoV-2 sarbecovirus lineage 378 responsible for the COVID-19 pandemic. Nature Microbiology 1–10 (2020) 379 doi:10.1038/s41564-020-0771-4. 380 26. Li, X. et al. Emergence of SARS-CoV-2 through recombination and strong purifying 381 selection. Science Advances 6, eabb9153 (2020). 382 27. Bernstein, F. C. et al. The Protein Data Bank: a computer-based archival file for 383 macromolecular structures. Journal of molecular biology 112, 535–42 (1977). 384 28. Adri C. T. van Duin, ‖, †, Siddharth Dasgupta, ‡, and Francois Lorant, \S & William 385 A. Goddard III*, ‡. ReaxFF: A Reactive Force Field for Hydrocarbons. (2001) 386 doi:10.1021/JP004368U. 387 29. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. Journal 388 of molecular graphics 14, 33–8, 27–8 (1996). 389 30. Tischer, B. K., Smith, G. A. & Osterrieder, N. En passant mutagenesis: a two step 390 markerless red recombination system. Methods Mol Biol 634, 421–430 (2010). 391 31. Koepke, L. et al. An improved method for high-throughput quantification of 392 autophagy in mammalian cells. Scientific Reports 10, 1–20 (2020). 393 32. Hohwieler, M. et al. ‘Miniguts’ from plucked human hair meet Crohn’s disease. Z 394 Gastroenterol 54, 748–759 (2016). 395 33. Larkin, M. A. et al. Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947– 396 2948 (2007). 397 34. Wagih, O. ggseqlogo: a versatile R package for drawing sequence logos. 398 Bioinformatics 33, 3645–3647 (2017). 399

400 Acknowledgments 15

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

401 We thank Kerstin Regensburger, Regina Burger, Jana-Romana Fischer, Birgit Ott, Martha

402 Meyer, Nicole Schrott and Daniela Krnavek for technical assistance. The ACE2 vector and the

403 SARS-CoV-2 S-HA plasmid were kindly provided by Shinji Makino and Stefan Pöhlmann,

404 and bat cells by Marcel A. Müller. F.Z., C.P.B., J.K. and L.K. are part of the International

405 Graduate school for Molecular Medicine (IGradU), Ulm. This study was supported by DFG

406 grants to F.K. (CRC 1279, SPP 1923), K-K.C. (Co260/6-1 Neuro-COVID), T.J. (CRC1279),

407 A.K. (KL 2544/8-1, KL 2544/5-1,7-1 and the 'Heisenberg-Programm' KL 2544/6-1) and

408 K.M.J.S. (CRC1279, SP1600/6-1). A.E. is funded by the State of Bavaria "BAY-VoC" and

409 "Coronaforschung". F.K., K.M.J.S. and A.E. were supported by the BMBF (Restrict SARS-

410 CoV-2, IMMUNOMOD and 01KI20172A SENSE-CoV2).

411 Author Contributions

412 F.Z. performed most experiments. D.S., M.V., Q.X. and L.K. performed western blots and

413 interaction assays. J.K., S.H. and A.K. generated and provided gut organoids. C.J. and T.J.

414 performed molecular modelling analyses. K.-K.C. provided pseudotypes and reagents. F.Z.,

415 D.S., K.M.J.S. and F.K. conceived the study, planned experiments and wrote the manuscript.

416 All authors reviewed and approved the manuscript.

417 Competing interests

418 The authors declare no competing interests.

419 Materials & Correspondence

420 Further information and requests for resources and reagents should be directed to and will be

421 fulfilled by Frank Kirchhoff ([email protected]).

422 423 FIGURES

424 Fig. 1: Modelling of the interaction of Coronavirus Spike residue 403 with human ACE2.

425 a, Schematic representation of the SARS-CoV-2 S protein (top panel), domains are indicated

426 in different colors. Receptor binding domain (RBD), light green. Receptor binding motif 16

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

427 (RBM), dark green. Transmembrane domain (TM), orange. R403, pink. S1/S2 and S2’

428 cleavage sites are indicated. Sequence alignment of SARS-CoV-2, SARS-CoV-1, Pan-CoV

429 and RaTG13 Spike RBD (bottom panel). Sequence conservation is indicated. purple arrows

430 denote important residues for ACE2 binding. b, Reactive force field simulation of SARS-CoV-

431 2 Spike in complex with human ACE2 (PDB: 7KNB) (left panel) and focus on position 403 in

432 SARS-CoV-2 S (R) or RaTG13 S (T) or respective exchange mutants at position 403 (right

433 panel). c, Exemplary energy curve of the reactive molecular dynamics simulation for SARS-

434 CoV-2 S and SARS-CoV-2 S R403T (top panel) and RaTG13 and RaTG13 T430R spike with

435 human ACE2 (bottom panel).

436 Fig. 2: R403 in Spike is crucial to use ACE2 as an entry receptor. a, Binary images of

437 CaCo2 cells transduced with VSVΔG-GFP pseudotyped with SARS-CoV-2, RaTG13 or

438 indicated mutant S. Successful infection events (=GFP positive cells) displayed as black dots.

439 Scale bar, 1.5mm. b, Automatic quantification of infection events by counting GFP positive

440 cells. n=3 (biological replicates) ± SEM. c, Bright field and fluorescence microscopy (GFP)

441 images of HEK293T cells transfected with SCoV-2ΔS bacmid, SCoV2-N, ACE2, T7

442 polymerase and indicated Spike variants. Scale bar, 125μm. d, Quantification of Gaussia

443 luciferase activity in the supernatant of HEK293T cells expressing SCoV-2ΔS-Gaussia

444 bacmids as described in (c). n=3 (biological replicates) ± SEM. P values are indicated

445 (student’s t test).

446 Fig. 3: T403R allows RaTG13 S to mediate infection of human gut organoids. a, Bright

447 field and fluorescence microscopy (GFP) images of hPSC derived gut organoids infected with

448 VSVΔG-GFP (green) pseudotyped with SARS-CoV-2, RaTG13 or indicated mutant S (300 µl,

449 2 h). Scale bar, 250μm. b, Quantification of the percentage of GFP-positive cells of (a). n=3

450 (biological replicates) ± SEM. P values are indicated (student’s t test).

17

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

451 Fig. 4: SARS-CoV-2 S and T403R RaTG13 S allow entry with human but not bat ACE2.

452 a, HEK293T cells expressing indicated ACE2 (Human ACE2 or Rhinolophus affinis ACE2)

453 constructs or b, Tb 1 Lu, Tadarida brasiliensis derived lung epithelial and Ri 1 Lu huACE2

454 Rhinolophus affinis derived lung epithelial cells expressing human ACE2 were infected with

455 VSVΔG-GFP pseudotyped with SARS-CoV-2, RaTG13 or indicated mutant S. Quantification

456 by automatic counting of GFP positive cells. n=3 (biological replicates) ± SEM. P values are

457 indicated (student’s t test).

458

459 Extended Figure legends

460 Extended Data Fig. 1: T403R RaTG13 S allows ACE2 dependent cell fusion. Exemplary

461 fluorescence microscopy images of HEK293T cells expressing SCoV2 S, RaTG13 S or the

462 indicated mutant, Human ACE2 and GFP (green). Insets are indicated by white boxes. Scale

463 bar, 125μm.

464 Extended Data Fig. 2: Incorporation of Spike variants in VSV pseudoparticles. a,

465 Exemplary immunoblots of whole cells lysates (WCLs) and supernatants of HEK293T cells

466 expressing SCoV2 S, RaTG13 S or the indicated mutant that were infected with VSVΔG-GFP.

467 Blots were stained with anti-SARS-CoV-2 S, anti-GAPDH and anti-VSV-M. b, Quantification

468 of Spike expression. n=3 (biological replicates) ± SEM. P values are indicated (student’s t test).

469 Extended Data Fig. 3: Processing of Spike proteins by ACE2 expression. a, Exemplary

470 immunoblots of WCLs of HEK293T cells expressing SARS-CoV-2 S, RaTG13 S or the

471 indicated mutant coexpressing Human ACE2 or empty vector construct. The blots were stained

472 with anti-SARS-CoV-2 S, anti-GAPDH, anti-ACE2 and anti-VSV-M.

473 Extended Data Fig. 4: SARS-CoV-2 entry is independent of α5β5 integrin. a, Automated

474 quantification by GFP fluorescence of Caco-2 cells preincubated with indicated amounts of

18

bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.

475 α5β5 integrin Inhibitor ATN-161 and infected with VSVΔG-GFP pseudotyped with SARS-

476 CoV-2, RaTG13 T403R mutant or RaTG13 S. n=3 (biological replicates) ± SEM. b,

477 Quantification of viral RNA copies in the supernatant of Calu-3 cells preincubated with

478 indicated amounts of ATN-161 and infected SARS-CoV-2 (MOI 0.05, 6 h). n=3 (biological

479 replicates) ± SEM. P values are indicated (student’s t test).

480 Extended Data Fig. 5: T403R allows RaTG13 S to mediate infection of human intestinal

481 organoids. a, Bright field and fluorescence microscopy (GFP) images of hPSC derived gut

482 organoids infected with equal amounts of VSVΔG-GFP (green) pseudotyped with SARS-CoV-

483 2, RaTG13 or indicated mutant S (2 h). Scale bar, 250μm. b, Exemplary gating strategy of flow

484 cytometry-based analysis of GFP-positive cells of (a). c, Quantification and exemplary gating.

485 n=3 (biological replicates) ± SEM. P values are indicated (student’s t test).

486 Extended Data Fig. 6: Bat ACE2 can be used for entry by SARS-CoV-2 Spike.

487 Quantification of GFP positive HEK293T cells expressing indicated ACE2 variants

488 (Rhinolophus macrotis ACE2 or Rhinolophus rhodesiae ACE2) infected with VSVΔG-GFP

489 pseudotyped with SARS-CoV-2, RaTG13 or indicated mutant S. n=3 (biological replicates) ±

490 SEM. P values are indicated (student’s t test).

491 Extended Data Fig. 7: Conservation of the RGD motif in bat Coronavirus Spike proteins.

492 a, Sequence logo of the alignment of 137 different bat Coronavirus Spike RBD sequences. The

493 RGD motif is highlighted by a red box. b, Primary sequence alignment of selected bat

494 coronaviruses, human coronaviruses and SARS-CoV-2 strains. The RGD motif is highlighted

495 in bold.

19 bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Fig.1 Zech et al. a RBD S1/S2 S2' N RBM TM C 1 319 437 517 541 685 815 1273 R403

SARS-CoV-2: NVYADSFVIRGDEVRQIAPGQTGKIADYNYKLPDDFTGCVIAWNSNNLDSKVGGNYNYLYRLFRKSNLKPFERDISTEIYQAGSTPCNGVKGFNCYFPLQSYGFQPTNGVGY 505 SARS-CoV-1: NVYADSFVVKGDDVRQIAPGQTGVIADYNYKLPDDFMGCVLAWNTRNIDATSTGNYNYKYRYLRHGKLRPFERDISNVPFSPDGKPCTP-PALNCYWPLNDYGFYTTTGIGY 491 Pan-CoV: NVYADSFVVKGDEVRQIAPGQTGVIADYNYKLPDDFTGCVIAWNSVKQDALTGGNYGYLYRLFRKSKLKPFERDISTEIYQAGSTPCNGQVGLNCYYPLERYGFHPTTGVNY 505 RaTG13: NVYADSFVITGDEVRQIAPGQTGKIADYNYKLPDDFTGCVIAWNSKHIDAKEGGNFNYLYRLFRKANLKPFERDISTEIYQAGSKPCNGQTGLNCYYPLYRYGFYPTDGVGH 505 ********: **:********** ************ ***:***: : *: **:.* ** :*:.:*:*******. :. ...**. .:***:** *** * *:.: b c 0 host cell SCoV-2 S SCoV-2 R403T S -5

E37 ] ACE2 V e [

E37 R403 E -10

T403 -15

SCoV 2 S 0 50 100 150 0 RaTG13 S E37 RaTG13 T403R S ACE2 E37 -5 ] V

R403 e [

E -10 T403

SCoV-2 RaTG13 S -15 0 50 100 150 t [ps] bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Fig. 2 Zech et al.

0.0033 a SCoV-2 S b 10000 vector SCoV-2 S T403R 0.0001 0.0001 1000

GFP 100 [cells/well] RaTG13 S RaTG13 S 10 RaTG13 S T403R T403A Pseudotype infection

1 GFP WT WT vector T403A R403T T403R SCoV-2 RaTG13 S S SCoV-2 S c d 15 vector SCoV-2 S R403T 0.0002

10 Brightfield

0.0008 0.0002 RaTG13 S RaTG13 S 5 RaTG13 S T403R T403A Gaussia [RLU/s x 10000] 0 WT WT Brightfield vector T403A R403T T403R SCoV-2 RaTG13 S S bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Fig. 3 Zech et al.

a SCoV-2 S RaTG13 S RaTG13 S b vector SCoV-2 S R403T RaTG13 S T403R T403A 100

GFP 0.031 0.029 50

Pseudotype infection [%] 0 WT WT Brightfield vector T403A R403T T403R SCoV-2 RaTG13 S S bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Fig. 4 Zech et al.

a Hu ACE2 Pu ACE2 b TB1 Lu1 Ri Lu1 huACE2 3×105 15000 0.0005 0.0009 200 200 <0.0001 <0.0001 0.0003 0.068 <0.0001 150 150 2×105 10000

100 100

1×105 5000 50 50

Pseudotype infection [cell/well] 0 0 0 0 WT WT WT WT WT WT WT WT vector vector vector vector T403A T403A T403A T403A R403T R403T R403T R403T T403R T403R T403R T403R SCoV-2 RaTG13 SCoV-2 RaTG13 SCoV-2 RaTG13 SCoV-2 RaTG13 S S S S S S S S bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Extended Data Fig. 1 Zech et al.

ACE2 SCoV2 S + ACE2 SCoV2 S R403T + ACE2 GFP

RaTG13 S + ACE2 RaTG13 T403R S + ACE2 RaTG13 T403A S + ACE2 GFP Extended DataFig.2 a bioRxiv preprint (which wasnotcertifiedbypeerreview)istheauthor/funder,whohasgrantedbioRxivalicensetodisplaypreprintinperpetuity.It cells supernatants 165 kDa- 165 kDa- 37 kDa- 60 kDa- 72 kDa- 60 kDa- 72 kDa- 93 kDa- 93 kDa- 37 kDa- [kDa]

vector doi: SCoV-2 WT S https://doi.org/10.1101/2021.05.31.446386 R403T

WT RaTG13

T403R S T403A made availableundera GAPDH VSV M S SCoV-2 GAPDH VSV M S SCoV-2 b normalized band intensity Spike/GAPDH 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 CC-BY-NC-ND 4.0Internationallicense SCoV-2 ;

S WT this versionpostedMay31,2021. R403T cells RaTG13 WT S T403R T403A SCoV-2 0.018

S WT supernatants R403T RaTG13 WT S T403R . T403A The copyrightholderforthispreprint Zech etal. bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Extended Data Fig. 3 Zech et al.

SCoV-2 S RaTG13 S WT WT WT WT T403R T403R T403A T403A R403T R403T vector - + - + - + - + - + - + ACE2 [kDa] - FL 165 kDa-

93 kDa- - S2* SCoV-2 Spike - S2 50 kDa-

120 kDa- ACE2

27 kDa- VSV M

36 kDa- GAPDH bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Extended Data Fig. 4 Zech et al. a b 1000 1000000 SARS CoV 2 S 800

600 100000 400 RaTG13 S T403R

200 GFP positive [cells/well] GFP SCoV2 N RNA [copies/mL] SCoV2 N RNA 0 RaTG13 S 10000

ATN-161 [μM] mock ATN-161 [μM] bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Extended Data Fig. 5 Zech et al.

a vector SCoV-2 S SCoV-2 S R403T RaTG13 S RaTG13 S T403R RaTG13 S T403A

Brightfield GFP Brightfield GFP Brightfield GFP Brightfield GFP Brightfield GFP Brightfield GFP

b gating mock SCoV-2 S SCoV-2 S R403T 0.019 % 1.29 % 1.29 % SSC-A

RaTG13 S RaTG3 S T403A SCoV-2 S R403T 0.039 % 0.54 % 0.042 % FSC-H

FSC-A GFP GFP GFP bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Extended Data Fig. 6 Zech et al.

Ma ACE2 Rh ACE2 20000 600 <0.0001 <0.0001 15000 <0.0001 400 0.0012

10000

200 5000

Pseudotype infection [cells/well] 0 0 WT WT WT WT vector vector T403A T403A R403T R403T T403R T403R

SCoV-2 RaTG13 SCoV-2 RaTG13 S S S S bioRxiv preprint doi: https://doi.org/10.1101/2021.05.31.446386; this version posted May 31, 2021. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Extended Data Fig. 7 Zech et al.

a 3

2 Bits 1

0 Position

b bat coronaviruses

human coronaviruses

SCoV-2 Strains