Next Article in Journal
Catecholamines Promote Ovarian Cancer Progression through Secretion of CXC-Chemokines
Next Article in Special Issue
Immunosenescence and Cytomegalovirus: Exploring Their Connection in the Context of Aging, Health, and Disease
Previous Article in Journal
The Glycosaminoglycan Side Chains and Modular Core Proteins of Heparan Sulphate Proteoglycans and the Varied Ways They Provide Tissue Protection by Regulating Physiological Processes and Cellular Behaviour
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Brief Report

Re-Analysis of the Widely Used Recombinant Murine Cytomegalovirus MCMV-m157luc Derived from the Bacmid pSM3fr Confirms Its Hybrid Nature

Institute for Clinical and Molecular Virology, Friedrich-Alexander-Universität Erlangen-Nürnberg, 91054 Erlangen, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(18), 14102; https://doi.org/10.3390/ijms241814102
Submission received: 1 August 2023 / Revised: 8 September 2023 / Accepted: 13 September 2023 / Published: 14 September 2023
(This article belongs to the Special Issue The Research about Cytomegalovirus (CMV))

Abstract

:
Murine cytomegalovirus (MCMV), and, in particular, recombinant virus derived from MCMV-bacmid pSM3fr, is widely used as the small animal infection model for human cytomegalovirus (HCMV). We sequenced the complete genomes of MCMV strains and recombinants for quality control. However, we noticed deviances from the deposited reference sequences of MCMV-bacmid pSM3fr. This prompted us to re-analyze pSM3fr and reannotate the reference sequence, as well as that for the commonly used MCMV-m157luc reporter virus. A correct reference sequence for this frequently used pSM3fr, containing a repaired version of m129 (MCK-2) and the luciferase gene instead of ORF m157, was constructed. The new reference also contains the original bacmid sequence, and it has a hybrid origin from MCMV strains Smith and K181.

1. Introduction

Human cytomegalovirus (HCMV) is a widely distributed pathogen, with seroprevalence of the lifelong infection varying from 40% to more than 90%. HCMV is the most frequent congenital pathogen after infection in utero and a significant clinical problem in immunocompromised hosts, causing severe symptoms like pneumonia, retinitis, and hepatitis [1]. Research on HCMV is complicated by the strict tropism of the virus for human cells and tissue, which makes it necessary to use surrogates like related betaherpesviruses from model animals [2,3,4], or more recently, complicated xenograft models like those in immune-deficient mice [5]. Murine cytomegalovirus (MCMV) has a high sequence similarity to HCMV and, not least due to availability of transgenic hosts, is widely used in small animals as an infection model for HCMV. The route of infection of MCMV was analyzed by utilizing reporter viruses [6,7], showing that MCMV spreads from the initial nasal entrance site to secondary vascular sites (e.g., spleen) and subsequently to tertiary sites like the salivary glands. Furthermore, it was demonstrated that the primary target after intra-footpad infection is the popliteal lymph node, which is neither host- nor virus-strain-specific. Intra-footpad infection can thus simulate peripheral infection [6]. Experiments analyzing the natural oral and nasal route of infection show that, in this case, the olfactory neurons are the primary target cells. However, deeper infection of the lower respiratory tract is also possible. After nasal infection, the tertiary virus spreads to the salivary glands and therein establishes persistence [7]. The replication of CMV usually occurs in differentiated myeloid cells [8]. Typical laboratory strains of MCMV are the strains Smith and K181. The genomes of these strains show a very high sequence similarity to each other (nucleic acid identity > 99%). However, they are highly variable with respect to “private” gene families like m02 and m145, located near the opposite termini of the genome. These play an important role in viral antigen presentation and, therefore, immune evasion ([9], see also Figure 1). The MCMV Smith and K181 strains differ in their infection dynamics in vitro and in vivo. In vitro, K181 leads to lower viral titers and smaller plaques in comparison to Smith, while the titers of K181 in vivo are higher in salivary glands [10,11]. In general, there are distinct genetic differences in the CMV genomes infecting different hosts, and also between MCMV strains, which are often found in the area of surface proteins, suggesting host adaption and immune evasion [12,13].
Genetic modification of viral genomes with traditional methods is difficult and laborious. The first bacterial artificial chromosome (bacmid) containing a near-complete sequence of a herpesvirus, MCMV strain Smith by Messerle and colleagues [14], represented a major breakthrough. However, extensive analysis of this bacmid pSM3 detected a gap comprising the open reading frames (ORFs) m151-m158. The loss of this region might have occurred to reduce the oversized nature of the viral genome due to insertion of the bacmid sequence via recombination. The missing region was later reinserted using homologous recombination with cloned MCMV DNA sequences [15] derived from plasmids initially designated as strain Smith but later identified as being derived from MCMV strain K181 [16,17]. Therefore, the resulting bacmid pSM3fr represents a hybrid consisting of the m150 to m158 genes belonging to the K181 strain within the genome of strain Smith. Thereafter, several groups observed an impaired replication of the pSM3fr-derived viruses in salivary glands; this was finally shown to be attributable to a frameshift mutation in the MCK-2 (MCMV chemokine homologue) gene, encoded in the ORF m129, leading to impaired replication in salivary glands. Repair of the m129 frameshift in pSM3fr-MCK-2fl (full length) restored salivary gland replication [18]. Interestingly, a further gene, m155, was initially implicated in the deficiency in salivary gland replication of the pSM3fr-derived viruses [19]. Furthermore, the pSM3fr bacmid was used as a template for generation of various reporter viruses, e.g., MCMV-m157luc, with luciferase instead of m157. The M157 glycoprotein was identified as a ligand for the natural killer (NK) cell activation receptor Ly49H [20]. Interestingly, this interaction between a viral protein and a host cell receptor was shown to be the opposite of immune evasion in the scope of co-infection and mutation studies [21,22]. Therefore, the deletion of the m157 gene with subsequent replacement by a luciferase gene does not impair the fitness of the resulting viruses during infection but sensitizes C57BL6 mice due to reduced NK cell defense. In other mouse strains, which are Ly49H-negative, M157 might act as ligand for different receptors. The original pSM3fr bacmid and derivatives like MCMV-m157luc were used for multiple studies (e.g., [23,24]); MCMV-m157luc was also independently repaired in the m129 gene encoding MCK-2 [25]. In the scope of another project from our lab, we sequenced the full bacmids and viruses derived thereof, and we noted numerous divergences from the deposited m129-repaired pSM3fr-MCK-2fl reference sequence in GenBank (Acc. No. KY348373), which prompted us to reassess the history of this widely used reagent. During our effort to reconstruct the history of this important reagent, we generated new reference sequences for the original pSM3fr and the widely used luciferase-containing bacmids.

2. Results and Discussion

In the scope of generating modified MCMV bacmids derived from the widely used pSM3fr bacmid, we performed full-genome sequencing of the constructed DNA via next-generation sequencing (NGS) to verify the correct assembly and search for potential mutations. Since we detected a high number of unexpected mismatches in specific regions, we also sequenced the original pSM3fr bacmid [14,17].
The sequences generated from the original pSM3fr bacmid were mapped to the MCMV strain Smith (GenBank Acc. No.: OP429142) and the deposited pSM3fr-MCK-2fl reference sequence (GenBank Acc. No.: KY348373) with average coverages of 160 and 162, respectively. Interestingly, for both mappings, the variant detection showed a heavily mutated region in the open reading frames (ORFs) m150 and m151, which encode type 1 membrane proteins of the m145 family (Figure 2A). In total, for m150, we detected 19 SNPs that lead to 12 amino acid changes, and for m151, 51 SNPs that lead to 32 amino acid changes. The alignment of the consensus sequence of m150 and m151 with the same region of different MCMV strains revealed that these ORFs match the strain K181, while the rest of the sequences match the Smith strain (Figure 2B,C). These regions are, therefore, part of the reinserted ORFs to complete the MCMV genome within the bacmid, in order to repair an initially loss of m151-m158 during cloning [17]. This divergence was expected for the comparison with the Smith strain, but not for the deposited pSM3fr-MCK-2fl reference sequence [18]. Moreover, this reference sequence does not contain annotations for the bacmid sequences, which were found in our mapping subsequent to ORF m158, flanked by repetitive elements for recombination (Figure 3A). Lodha et al. recently studied the transcriptional profile of MCMV, including noncoding RNA of the pSM3fr-derived virus; in this context, they also adapted the deposited reference sequence KY348373 [26]. However, this sequence does not contain the corrections in the region m150 to m158. We constructed a new reference sequence for the hybrid pSM3fr bacmid, which contains the regions originating from K181 at the position of the ORFs m150 to m158, the frameshift mutation in m129 (encoding for MCMV chemokine homologue MCK-2 [18]), as well as the bacmid-specific regions between m158 and m159 (Figure 3B). Further variants that were detected after mapping our pSM3fr libraries to the deposited reference sequence (KY348373) are listed in Table 1. Of note, this earlier deposited pSM3fr MCK2-fl reference sequence was obviously corrected for the m129 frameshift, but not for any of the further variant positions with respect to strain Smith described in [18].
Another derivative of the widely used bacmid pSM3fr derived from MCMV contains a repaired version of the MCK-2 gene (m129) [18] and a HCMV IE promoter-driven luciferase cassette replacing ORF m157 (MCMV-m157luc) [23,24]. For future analysis, we also generated an annotated reference sequence for this bacmid derived from respective consensus sequences and confirmed it via resequencing (Figure 3C).
Furthermore, we sequenced two different stocks of low-passage viruses reconstituted from the MCMV-m157luc bacmid with an average coverage of 330. The resulting reads were then mapped to our newly generated reference sequence for MCMV-m157luc (Figure 4) and reached average coverages of 256 and 2523, respectively. Interestingly, both mappings show gaps within the area encoding the bacmid components. This indicates that loss of genomic regions unimportant for viral replication is not infrequent, most probably due to the large size of the MCMV genome, which is already at the edge of the nucleocapsid packaging capacity [17,27]. This probably was also the cause of the initially observed deletion of m151 to m158 in the original bacmid [14]. Since the genes encoded in the bacmid areas are dispensable for viral replication and infection, we suggest that viruses that lose these parts—instead of parts of the original viral genome—have replication advantages over viruses with the full bacmid. This is supported by the fact that we observe partial bacmid loss in two different viral stocks at different positions, but no gene loss or extensive gaps in other parts of the genome. Nevertheless, the tendency of gene loss and further genetic alterations during passaging demonstrates the usefulness of regular full-genome sequencing of both, newly generated viral stocks as well as passaged viruses. It also suggests that one should more regularly make use of the two loxP sites flanking the bacmid sequences; by passaging the virus through Cre recombinase-expressing cells, better comparable viruses with a homogenous deletion of the bacmid and a better packable genome size may be obtained. However, the loxP sites are not at the very ends of the bacmid, leaving residual foreign sequences. Alternatively, the vector cassette could be moved to the end of the genome or placed within an essential gene to increase selection pressure [28]. Since HCMV has a similar genome size to MCMV, we would assume similar parameters and restrictions apply for bacmid cloning and rescue. This indicates that particular attention is required in the characterization of recombinant herpesviruses, like marker rescue, careful restriction analysis and resequencing of full recombinant genomes.
Furthermore, since HCMV has a similar genome size to MCMV, we would expect a tendency of gene loss in bacmid systems as well and would recommend full-genome sequencing of bacmid-derived HMCV.
In conclusion, we provide the CMV research community with new and corrected references for the extensively used pSM3fr bacmid (accession: ERZ20801763) and the MCK-2-repaired version, MCMV-m157luc reporter virus, containing a luciferase instead of ORF m157 (accession: ERZ20801746).

3. Material and Methods

3.1. Bacterial Culture and DNA Preparation

Escherichia coli DH10B-derived EL250 [29] or GS1783 [30] carrying bacmid DNA was cultured overnight in LB medium containing 15 µg/mL chloramphenicol at 32 °C and 200 rpm. Bacmid DNA was purified using the PureLinkTM HiPure Plasmid Maxiprep Kit of Invitrogen, Thermo Fisher Scientific, Waltham, MA, USA (K210006) according to the manufacturer’s instructions.

3.2. Library Preparation and Sequencing

Purified bacmid DNA was obtained via standard precipitation protocols. Viral DNA was extracted utilizing a Qiagen (Venlo, The Netherlands) EZ-1 instrument. For the library preparation for next-generation sequencing, 500 ng of bacmid DNA or extracted viral DNA was fragmented and adapters and indices were ligated using the NEBNext® Ultra™ II FS DNA Library Prep Kit from Illumina, San Diego, CA, USA (E7805), according to manufacturer’s instructions, with a fragmentation time of 15 min.
Sequencing was performed via paired-end sequencing utilizing the MiSeq Reagent Kit v3 (150 cycles) on a MiSeq™ Instrument (Illumina, San Diego, CA, USA). Sequence analysis was conducted using CLC Genomics Workbenches 22 and 23 (Qiagen Aarhus A/S, Denmark). Raw reads were trimmed for quality (limit 0.05, Mott trimming algorithm), adapters, and ambiguities. The trimmed reads were mapped with different stringencies, testing for the best distribution. Duplicate reads were removed from the mapping. Variants from the reference sequence were detected, only considering regions with coverages above 50 and a frequency of the variant in the reads of more than 80%.

3.3. Production and Purification of MCMV Stocks

In order to generate high-titer stocks of MCMV, murine embryonic fibroblast (MEF) cells were seeded 1:2 into twelve T175 cell culture flasks. On the next day, purified MCMV virions were thawed and 1 × 105 virions were added to 25 mL of complete DMEM. After extensive vortexing, the MEF supernatant of each flask was exchanged for 25 mL of complete DMEM containing MCMV virions and the cells were incubated for 1 week under cell culture conditions. Cell culture supernatants were centrifuged (1500× g, 20 min, 4 °C) to remove cell debris. The supernatants were ultracentrifuged (33,000× g, 3 h, 4 °C). The supernatant was discarded; then, the pellets were resuspended in a few microliters of residual media, pooled, and dispersed using a syringe. Following that, 1 mL of concentrated virions was carefully loaded on a 9 mL cushion of 15% sucrose in virus suspension buffer (VSP, 50 mM Tris-HCl (pH 7.8), 12 mM KCl, 5 mM EDTA) and purified via ultracentrifugation (70,000× g, 1 h, 4 °C). The sucrose cushion was discarded; then, the pellets were washed in 10 mL of PBS and repelleted (70,000× g, 1 h, 4 °C). PBS was discarded and 500 µL of VSP was added and incubated on ice overnight. The next day, pellets were resuspended, pooled, dispersed using a syringe, and stored at −80 °C in 40 µL aliquots.

3.4. Sequences

Reference sequences were deposited in the European Nucleotide Archive (https://www.ebi.ac.uk/ena/browser/home, accessed on 31 July 2023). The accession number for the original pSM3fr bacmid is ERZ20801763, while the accession number for the MCK-2 repaired version with a luciferase instead of ORF m157 is ERZ20801746.

Author Contributions

Conceptualization, A.C. and A.E.; methodology, A.C., C.B., S.K. and L.B.; software, A.C. and A.E.; validation, A.C. and A.E.; formal analysis, A.C. and A.E.; investigation, A.C.; resources, A.C. and A.E.; data curation, A.C.; writing—original draft preparation, A.C. and A.E.; writing—review and editing, A.C., C.B., S.K., L.B. and A.E.; visualization, A.C. and A.E.; supervision, A.C. and A.E.; project administration, A.E.; funding acquisition, A.E. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by DFG 401821119/GRK 2504.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The generated reference sequences were deposited in the European Nucleotide Archive (https://www.ebi.ac.uk/ena/browser/home, accessed on 31 July 2023). The accession number for the original pSM3fr bacmid is ERZ20801763, while the accession number for the MCK-2 repaired version with a luciferase instead of ORF m157 is ERZ20801746. All other data are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kenneson, A.; Cannon, M.J. Review and meta-analysis of the epidemiology of congenital cytomegalovirus (CMV) infection. Rev. Med. Virol. 2007, 17, 253–276. [Google Scholar] [CrossRef] [PubMed]
  2. Fisher, M.A.; Lloyd, M.L. A Review of Murine Cytomegalovirus as a Model for Human Cytomegalovirus Disease-Do Mice Lie? Int. J. Mol. Sci. 2020, 22, 214. [Google Scholar] [CrossRef] [PubMed]
  3. Itell, H.L.; Kaur, A.; Deere, J.D.; Barry, P.A.; Permar, S.R. Rhesus monkeys for a nonhuman primate model of cytomegalovirus infections. Curr. Opin. Virol. 2017, 25, 126–133. [Google Scholar] [CrossRef] [PubMed]
  4. Roark, H.K.; Jenks, J.A.; Permar, S.R.; Schleiss, M.R. Animal Models of Congenital Cytomegalovirus Transmission: Implications for Vaccine Development. J. Infect. Dis. 2020, 221, S60–S73. [Google Scholar] [CrossRef]
  5. Crawford, L.D.; Streblow, D.N.; Hakki, M.; Nelson, J.A.; Caposio, P. Humanized mouse models of human cytomegalovirus infection. Curr. Opin. Virol. 2015, 13, 86–92. [Google Scholar] [CrossRef]
  6. Farrell, H.E.; Davis-Poynter, N.; Bruce, K.; Lawler, C.; Dolken, L.; Mach, M.; Stevenson, P.G. Lymph Node Macrophages Restrict Murine Cytomegalovirus Dissemination. J. Virol. 2015, 89, 7147–7158. [Google Scholar] [CrossRef]
  7. Farrell, H.E.; Lawler, C.; Tan, C.S.; MacDonald, K.; Bruce, K.; Mach, M.; Davis-Poynter, N.; Stevenson, P.G. Murine Cytomegalovirus Exploits Olfaction to Enter New Hosts. mBio 2016, 7, e00251-16. [Google Scholar] [CrossRef]
  8. Sinclair, J.; Reeves, M. The intimate relationship between human cytomegalovirus and the dendritic cell lineage. Front. Microbiol. 2014, 5, 389. [Google Scholar] [CrossRef]
  9. Becker, S.; Fink, A.; Podlech, J.; Reddehase, M.J.; Lemmermann, N.A. Host-Adapted Gene Families Involved in Murine Cytomegalovirus Immune Evasion. Viruses 2022, 14, 128. [Google Scholar] [CrossRef]
  10. Hudson, J.B.; Walker, D.G.; Altamirano, M. Analysis in vitro of two biologically distinct strains of murine cytomegalovirus. Arch. Virol. 1988, 102, 289–295. [Google Scholar] [CrossRef]
  11. Misra, V.; Hudson, J.B. Minor Base Sequence Differences Between the Genomes of Two Strains of Murine Cytomeflalovirus Differin 9 in Virulence. Arch. Virol. 1980, 64, 1–8. [Google Scholar] [CrossRef] [PubMed]
  12. Cizkova, D.; Baird, S.J.E.; Tesikova, J.; Voigt, S.; Ludovit, D.; Pialek, J.; Gouy de Bellocq, J. Host subspecific viral strains in European house mice: Murine cytomegalovirus in the Eastern (Mus musculus musculus) and Western house mouse (Mus musculus domesticus). Virology 2018, 521, 92–98. [Google Scholar] [CrossRef] [PubMed]
  13. Smith, L.M.; McWhorter, A.R.; Shellam, G.R.; Redwood, A.J. The genome of murine cytomegalovirus is shaped by purifying selection and extensive recombination. Virology 2013, 435, 258–268. [Google Scholar] [CrossRef] [PubMed]
  14. Messerle, M.; Crnkovic, I.; Hammerschmidt, W.; Ziegler, H.; Koszinowski, U.H. Cloning and mutagenesis of a herpesvirus genome as an infectious bacterial artificial chromosome. Proc. Natl. Acad. Sci. USA 1997, 94, 14759–14763. [Google Scholar] [CrossRef]
  15. Mercer, J.A.; Marks, J.R.; Spector, D.H. Molecular Cloning and Restriction Endonuclease Mapping of the Murine Cytomegalovirus Genome (Smith Strain). Virology 1983, 129, 94–106. [Google Scholar] [CrossRef]
  16. Elliot, R.; Clark, C.; Jaquish, D.; Spector, D.H. Transcription Analysis and Sequence of the Putative Murine Cytomegalovirus DNA Polymerase Gene. Virology 1991, 185, 169–186. [Google Scholar] [CrossRef]
  17. Wagner, M.; Jonjic, S.; Koszinowski, U.H.; Messerle, M. Systematic excision of vector sequences from the BAC-cloned herpesvirus genome during virus reconstitution. J. Virol. 1999, 73, 7056–7060. [Google Scholar] [CrossRef]
  18. Jordan, S.; Krause, J.; Prager, A.; Mitrovic, M.; Jonjic, S.; Koszinowski, U.H.; Adler, B. Virus progeny of murine cytomegalovirus bacterial artificial chromosome pSM3fr show reduced growth in salivary Glands due to a fixed mutation of MCK-2. J. Virol. 2011, 85, 10346–10353. [Google Scholar] [CrossRef]
  19. Hasan, M.; Krmpotic, A.; Ruzsics, Z.; Bubic, I.; Lenac, T.; Halenius, A.; Loewendorf, A.; Messerle, M.; Hengel, H.; Jonjic, S.; et al. Selective Down-Regulation of the NKG2D Ligand H60 by Mouse Cytomegalovirus m155 Glycoprotein. J. Virol. 2005, 79, 2920–2930. [Google Scholar] [CrossRef]
  20. Smith, H.R.; Heusel, J.W.; Mehta, I.K.; Kim, S.; Dorner, B.G.; Naidenko, O.V.; Iizuka, K.; Furukawa, H.; Beckman, D.L.; Pingel, J.T.; et al. Recognition of a virus-encoded ligand by a natural killer cell activation receptor. Proc. Natl. Acad. Sci. USA 2002, 99, 8826–8831. [Google Scholar] [CrossRef]
  21. Voigt, V.; Forbes, C.A.; Tonkin, J.N.; Degli-Esposti, M.A.; Smith, H.R.; Yokoyama, W.M.; Scalzo, A.A. Murine cytomegalovirus m157 mutation and variation leads to immune evasion of natural killer cells. Proc. Natl. Acad. Sci. USA 2003, 100, 13483–13488. [Google Scholar] [CrossRef]
  22. McWhorter, A.R.; Smith, L.M.; Masters, L.L.; Chan, B.; Shellam, G.R.; Redwood, A.J. Natural killer cell dependent within-host competition arises during multiple MCMV infection: Consequences for viral transmission and evolution. PLoS Pathog. 2013, 9, e1003111. [Google Scholar] [CrossRef] [PubMed]
  23. Klenovsek, K.; Weisel, F.; Schneider, A.; Appelt, U.; Jonjic, S.; Messerle, M.; Bradel-Tretheway, B.; Winkler, T.H.; Mach, M. Protection from CMV infection in immunodeficient hosts by adoptive transfer of memory B cells. Blood 2007, 110, 3472–3479. [Google Scholar] [CrossRef] [PubMed]
  24. Bootz, A.; Karbach, A.; Spindler, J.; Kropff, B.; Reuter, N.; Sticht, H.; Winkler, T.H.; Britt, W.J.; Mach, M. Protective capacity of neutralizing and non-neutralizing antibodies against glycoprotein B of cytomegalovirus. PLoS Pathog. 2017, 13, e1006601. [Google Scholar] [CrossRef] [PubMed]
  25. Sell, S.; Dietz, M.; Schneider, A.; Holtappels, R.; Mach, M.; Winkler, T.H. Control of murine cytomegalovirus infection by gammadelta T cells. PLoS Pathog. 2015, 11, e1004481. [Google Scholar] [CrossRef]
  26. Lodha, M.; Muchsin, I.; Jurges, C.; Juranic Lisnic, V.; L’Hernault, A.; Rutkowski, A.J.; Prusty, B.K.; Grothey, A.; Milic, A.; Hennig, T.; et al. Decoding murine cytomegalovirus. PLoS Pathog. 2023, 19, e1010992. [Google Scholar] [CrossRef]
  27. Smith, L.M.; McWhorter, A.R.; Masters, L.L.; Shellam, G.R.; Redwood, A.J. Laboratory strains of murine cytomegalovirus are genetically similar to but phenotypically distinct from wild strains of virus. J. Virol. 2008, 82, 6689–6696. [Google Scholar] [CrossRef]
  28. Sampaio, K.L.; Weyell, A.; Subramanian, N.; Wu, Z.; Sinzger, C. A TB40/E-derived human cytomegalovirus genome with an intact US-gene region and a self-excisable BAC cassette for immunological research. Biotechniques 2017, 63, 205–214. [Google Scholar] [CrossRef]
  29. Lee, E.C.; Yu, D.; Martinez de Velasco, J.; Tessarollo, L.; Swing, D.A.; Court, D.L.; Jenkins, N.A.; Copeland, N.G. A highly efficient Escherichia coli-based chromosome engineering system adapted for recombinogenic targeting and subcloning of BAC DNA. Genomics 2001, 73, 56–65. [Google Scholar] [CrossRef]
  30. Tischer, B.K.; Smith, G.A.; Osterrieder, N. En passant mutagenesis: A two step markerless red recombination system. Methods Mol. Biol. 2010, 634, 421–430. [Google Scholar] [CrossRef]
Figure 1. Genome alignment of the MCMV strains Smith and K181. The reference sequences for the MCMV strains Smith (OP429142) and K181 (AM88612) from GenBank were aligned and sequence conservation is displayed. Black bars indicate lower conservation. The genome areas of m02-m06 as well as m145-m159 are enlarged to demonstrate high sequence diversity in these respective areas between the two strains.
Figure 1. Genome alignment of the MCMV strains Smith and K181. The reference sequences for the MCMV strains Smith (OP429142) and K181 (AM88612) from GenBank were aligned and sequence conservation is displayed. Black bars indicate lower conservation. The genome areas of m02-m06 as well as m145-m159 are enlarged to demonstrate high sequence diversity in these respective areas between the two strains.
Ijms 24 14102 g001
Figure 2. Comparison of ORFs m150 and m151 between MCMV strains Smith and K181. (A) Mapping of original pSM3fr reads against MCMV strain Smith and pSM3fr-MCK-2fl reference sequence and variant detection. Colored vertical lines indicate multiple nucleotide polymorphisms. (B) Section of an alignment of pSM3fr-MCK-2fl sequence of ORFs m151 and m152 as well as the respective regions of database reference sequences for MCMV strains Smith (OP429142, GU305914) and K181 (AM88612). Alignment was constructed using the Jukes–Cantor model. Black bars indicate lower conservation. (C) Maximum-likelihood tree generation with 1000 bootstrap repeats from the alignment of (B); numbers at bifurcation give percent of trees with the respective branching. Scale bars indicate the phylogenetic distance in percent of number of substitutions/changes per nucleotide.
Figure 2. Comparison of ORFs m150 and m151 between MCMV strains Smith and K181. (A) Mapping of original pSM3fr reads against MCMV strain Smith and pSM3fr-MCK-2fl reference sequence and variant detection. Colored vertical lines indicate multiple nucleotide polymorphisms. (B) Section of an alignment of pSM3fr-MCK-2fl sequence of ORFs m151 and m152 as well as the respective regions of database reference sequences for MCMV strains Smith (OP429142, GU305914) and K181 (AM88612). Alignment was constructed using the Jukes–Cantor model. Black bars indicate lower conservation. (C) Maximum-likelihood tree generation with 1000 bootstrap repeats from the alignment of (B); numbers at bifurcation give percent of trees with the respective branching. Scale bars indicate the phylogenetic distance in percent of number of substitutions/changes per nucleotide.
Ijms 24 14102 g002
Figure 3. Schematic representation of corrected reference sequences. (A) Mapping showing the bacmid-specific sequences within the references. (B) The corrected pSM3fr reference sequence, with the frameshift mutation in m129 and strain-K181-derived sequences in m151-m158. (C) The newly generated reference sequence for the repaired and modified pSM3fr-m157luc bacmid, containing the repaired version of m129 and a luciferase gene cassette instead of m157.
Figure 3. Schematic representation of corrected reference sequences. (A) Mapping showing the bacmid-specific sequences within the references. (B) The corrected pSM3fr reference sequence, with the frameshift mutation in m129 and strain-K181-derived sequences in m151-m158. (C) The newly generated reference sequence for the repaired and modified pSM3fr-m157luc bacmid, containing the repaired version of m129 and a luciferase gene cassette instead of m157.
Ijms 24 14102 g003
Figure 4. Mapping of reads obtained from low-passage MCMV viruses (m157-luc). Displayed are the mapped reads from two different viral stocks. The genome area spanning the bacmid components, which are flanked by the repetitive elements for recombination, is shown. In both cases, gaps (without specific read mapping or with strongly reduced specific read mapping) can be observed.
Figure 4. Mapping of reads obtained from low-passage MCMV viruses (m157-luc). Displayed are the mapped reads from two different viral stocks. The genome area spanning the bacmid components, which are flanked by the repetitive elements for recombination, is shown. In both cases, gaps (without specific read mapping or with strongly reduced specific read mapping) can be observed.
Ijms 24 14102 g004
Table 1. Detected variants of our pSM3fr mapping to the deposited pSM3fr reference sequence (KY348373). Only variants that are observed in above 80% of the reads are shown. ORF = open reading frame, fs = frameshift, ins = insertion, - = gap, SNV = single-nucleotide variant, MNV = multi-nucleotide variant, ? = unknown effect.
Table 1. Detected variants of our pSM3fr mapping to the deposited pSM3fr reference sequence (KY348373). Only variants that are observed in above 80% of the reads are shown. ORF = open reading frame, fs = frameshift, ins = insertion, - = gap, SNV = single-nucleotide variant, MNV = multi-nucleotide variant, ? = unknown effect.
ORFPositionTypeFrequency [%]Amino Acid ChangeIn
Reference
In
Mapping
Variants already described in [18] but not corrected in the sequence
m2020957Deletion97Q697fsG-
intergenic31973Insertion91none-G
m29, m29.136197Insertion99S297fs; D155fs-G
intergenic37259Insertion84none-C
M3138802Insertion87A509fs-G
m45.161917Insertion88A375fs-C
m5892088Deletion96N110fsA-
m5892358Deletion100P200fsC-
m142.1201402Insertion93G477fs-G
intergenic214373Deletion99noneT-
m156215809Insertion87Insertion of P between V56 and S57-CGG
intergenic219102Insertion90none-A
intergenic219308SNV98noneCA
intergenic219496SNV97noneGA
intergenic221004Insertion94none-G
intergenic221024Insertion87none-GG
intergenic226294Deletion98noneC-
m161228998Insertion85S187fs-A
Variants due to the hybrid nature of the bacmid (K181 from m150 to m158)
m150208207SNV96K195QTG
m150208211SNV93noneAC
m150208214SNV95noneCT
m150208235SNV94noneAG
m150208243MNV99Replacement of E182 and V183 by D and ICCTA
m150208250MNV98R180NACGGTT
m150208261SNV98V177LCA
m150208280SNV100noneGC
m150208283SNV98noneGA
m150208289SNV100noneTC
m150208500Insertion99W97fs-G
m150208502Deletion96W97fsC-
m150208503Insertion88L96fs-T
m150208505Deletion85L96fsA-
m150208508SNV88noneAG
m150208510SNV93H94YGA
m150208525SNV98D89NCT
m150208529MNV98K87TCTAG
m150208788Insertion90M1?-T
m151209113Insertion72E290 duplicate-TTC
m151209126Deletion98A287fsA-
m151209127Insertion97A286fs-T
m151209129SNV98noneAG
m151209138SNV99noneCT
m151209154MNV99T277VGTAC
m151209157SNV100A276VGA
m151209161SNV100A275TCT
m151209258SNV98noneTC
m151209262SNV96S241WGC
m151209273SNV99N237KAC
m151209282SNV99noneAG
m151209289SNV100V232AAG
m151209291SNV98noneAC
m151209294SNV98noneTC
m151209348SNV93noneTG
m151209354SNV96noneGA
m151209357SNV99noneTA
m151209359MNV97Replacement of A208 and T209 by V and ATTGCCA
m151209375SNV90noneTC
m151209379MNV97N202GTTCC
m151209383SNV99T201STA
m151209399MNV99S195FGGAA
m151209403MNV97V194TACCGTT
m151209406Insertion97S193fs-T
m151209408Deletion94S193fsC-
m151209414SNV96noneTC
m151209416SNV99K190ETC
m151209419SNV94T189ATC
m151209437MNV100L183IGCTA
m151209440SNV96T182STA
m151209446MNV98I180LTGAA
m151209452SNV100Q178KGT
m151209457SNV94R176KCT
m151209459SNV93noneGA
m151209461Deletion97R175fsG-
m151209462Insertion100R175fs-G
m151209467MNV100I173LTAGG
m151209471SNV100noneAT
m151209474SNV100noneAG
m151209576Insertion73P138fs-AC
m151209579Replacement92I135fsTAC
m151209583Replacement100L134fsAAG
m151209588MNV98A132EGGCT
m151209597SNV93noneGC
m151209603SNV99noneAG
m151209605Insertion94Y127fs-G
m151209607Deletion97Y126fsT-
m151209801SNV96noneGC
m151209815SNV97D57NCT
m151209818MNV98Y56HACGT
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cordsmeier, A.; Bednar, C.; Kübel, S.; Bauer, L.; Ensser, A. Re-Analysis of the Widely Used Recombinant Murine Cytomegalovirus MCMV-m157luc Derived from the Bacmid pSM3fr Confirms Its Hybrid Nature. Int. J. Mol. Sci. 2023, 24, 14102. https://doi.org/10.3390/ijms241814102

AMA Style

Cordsmeier A, Bednar C, Kübel S, Bauer L, Ensser A. Re-Analysis of the Widely Used Recombinant Murine Cytomegalovirus MCMV-m157luc Derived from the Bacmid pSM3fr Confirms Its Hybrid Nature. International Journal of Molecular Sciences. 2023; 24(18):14102. https://doi.org/10.3390/ijms241814102

Chicago/Turabian Style

Cordsmeier, Arne, Christopher Bednar, Sabrina Kübel, Larissa Bauer, and Armin Ensser. 2023. "Re-Analysis of the Widely Used Recombinant Murine Cytomegalovirus MCMV-m157luc Derived from the Bacmid pSM3fr Confirms Its Hybrid Nature" International Journal of Molecular Sciences 24, no. 18: 14102. https://doi.org/10.3390/ijms241814102

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop