World Scientific
  • Search
  •   
Skip main navigation

Cookies Notification

We use cookies on this site to enhance your user experience. By continuing to browse the site, you consent to the use of our cookies. Learn More
×
Our website is made possible by displaying certain online content using javascript.
In order to view the full content, please disable your ad blocker or whitelist our website www.worldscientific.com.

System Upgrade on Tue, Oct 25th, 2022 at 2am (EDT)

Existing users will be able to log into the site and access content. However, E-commerce and registration of new users may not be available for up to 12 hours.
For online purchase, please visit us again. Contact us at [email protected] for any enquiries.

Brain-wide N2cG compensation permits glycoprotein-deleted rabies virus to trace neural circuits across multiple synapses

    https://doi.org/10.1142/S1793545823400011Cited by:1 (Source: Crossref)

    Abstract

    Rabies-viruses-based retrograde tracers can spread across multiple synapses in a retrograde direction in the nervous system of rodents and primates, making them powerful tools for determining the structure and function of the complicated neural circuits of the brain. However, they have some limitations, such as posing high risks to human health and the inability to retrograde trans-synaptic label inputs from genetically-defined starter neurons. Here, we established a new retrograde trans-multi-synaptic tracing method through brain-wide rabies virus glycoprotein (RVG) compensation, followed by glycoprotein-deleted rabies virus (RV-ΔG) infection in specific brain regions. Furthermore, in combination with the avian tumor virus receptor A (TVA) controlled by a cell-type-specific promoter, we found that EnvA-pseudotyped RV-ΔG can mediate efficient retrograde trans-multi-synaptic transduction from cell-type-specific starter neurons. This study provides new alternative methods for neuroscience researchers to analyze the input neural networks of rodents and nonhuman primates.

    1. Introduction

    Analyzing the input and output pathways of the complex neural circuits in the brain is the basis of revealing the principles of the brain function and elucidating the mechanisms through which brain diseases occur.1,2 Retrograde tracers based on neurotropic viruses have been continuously developed to achieve structural imaging and functional manipulation of input networks in different model animals, providing valuable gene delivery tools for research in neuroscience and gene therapy.3,4,5,6,7

    Retrograde viral tracers include retrograde trans-synaptic viruses and retrograde non-trans-synaptic viruses. The commonly used retrograde viral tracers are adeno-associated virus (AAV),6,8 lentivirus (LV),9 canine adenovirus (CAV),10 rabies virus (RV),11 herpes simplex virus (HSV),12 and pseudorabies virus (PRV).13 Among these viruses, only the first three viruses cannot be retrogradely transmitted across synapses. The toxicity of retrograde non-trans-synaptic viruses is generally lower than that of trans-synaptic virus, and some of them are used as gene therapeutic vectors. However, the retrograde non-trans-synaptic viruses cannot be directly used to trace the whole-brain synaptic inputs from cell-type-specific starter cells, whereas the retrograde trans-synaptic viruses can. Retrograde non-trans-synaptic viruses are mainly used to study the structural and functional networks of projection neurons; moreover, they can be used together with retrograde trans-synaptic viruses, enabling trans-synaptic input tracing from specific subsets of neurons based on their projection and cell type.8,14

    RV and PRV Bartha are retrograde trans-multi-synaptic viral vectors that can be used to map multiple input networks in the peripheral and central nervous systems.15 One important feature of RV is that it can infect primates, whereas PRV Bartha cannot, so RV is a powerful viral tool for dissecting the complex neural connections within the nervous system, especially in nonhuman primates.16,17 In general, wild-type RV is deadly to both humans and other animals; thereby, extreme caution is required for its use: Researchers working with it need to be vaccinated.16 Because PRV Bartha does not infect humans, recombinant PRV Bartha has been preferentially used for the neural circuit analysis of the peripheral and central nervous systems in rodents.18 For the precise identification of special neural circuits, PRV-Ba2001, a Cre-dependent tracer that can be retrogradely transmitted from originally infected cell-type-specific neurons, was developed.19 However, strategies based on Cre-dependent DNA recombination for cell-type-specific labeling cannot be applied to RV, because RV is an RNA virus and has no DNA phase in the viral life cycle.16 In addition, RV and PRV exhibit differential infection tropism in the central nervous system of rodents.20 Therefore, RV-based retrograde multi-trans-synaptic technology needs to be developed.

    Several attenuated rabies viral vectors have been developed, such as the SAD-B1921,22 and CVS-N2c23 strains. Glycoprotein-deleted rabies virus (RV-ΔG) is deficient in replication and thus cannot propagate beyond the originally infected neurons, unless RV glycoprotein (RVG) is supplied.11 Based on this, a RV-ΔG-based retrograde trans-monosynaptic system was developed.21 CVS-N2c-ΔG exhibits a reduction in neuronal toxicity and a substantial enhancement in trans-synaptic neuronal transfer compared with the SAD-B19-ΔG vaccine strain.23,24 RV-ΔG pseudotyped with avian sarcoma and leukosis virus A protein (EnvA) can only infect cells that express the avian tumor virus receptor A (TVA); according to this, specific infection can be achieved.15,21

    In this study, through brain-wide N2cG compensation using AAV-PHP.eB, followed by CVS-N2c-ΔG infection in specific brain regions, we established a new retrograde trans-multi-synaptic tracing method. Furthermore, in combination with the avian TVA controlled by a cell-type-specific promoter, we achieved retrograde trans-multi-synaptic tracing from cell-type-specific starter neurons. Our study provides new alternative methods for neuroscience researchers to analyze the input neural networks of rodents and nonhuman primates.

    2. Materials and Methods

    2.1. Plasmid construction and viral vector preparation

    To construct pAAV-EF1α-N2cG-WPRE-pA, we amplified N2cG gene sequence by PCR, which we inserted into DIO-EF1a-FLEX-H2B-GFP-P2A-N2cG (Addgene plasmid #73476) digested by restriction enzymes SalI and EcoRI (New England Biolabs). pAAV-CaMKIIα-EGFP-T2A-TVA-WPRE-pA was provided as a gift from BrainCase Biotechnology Co., Ltd. (ShenZhen, China). We produced the AAV vectors, including rAAV-PHP.eB-EF1α-N2cG-WPRE-pA and rAAV9-CaMKIIα-EGFP-T2A-TVA-WPRE-pA, by triple-plasmid transient transfection in HEK293T cells, pUCmini-iCAP-PHP.eB (Addgene plasmid #103005) or pAAV2/9n (Addgene plasmid #112865) as the packaging plasmid and pAdDeltaF6 (Addgene plasmid #112867) as the adenovirus element helper plasmid. The virus purification and concentration procedures were performed as previously described.3,5,25 The titers of rAAV-PHP.eB-EF1α-N2cG-WPRE-pA and rAAV9-CaMKIIα-EGFP-T2A-TVA-WPRE-pA were 1.36×1013 viral genomes/mL (VG/mL) and 2.0×1012VG/mL, respectively. We prepared B19G-CVS-N2c-ΔG-EGFP and EnvA-CVS-N2c-ΔG-tdTomato according to a previously reported method.24 The titers of B19G-CVS-N2c-ΔG-EGFP and EnvA-CVS-N2c-ΔG-tdTomato were 3×108 infectious units per mL(IU/mL) and 1×108IU/mL, respectively.

    2.2. Research animals

    All surgical and experimental procedures were approved by the Animal Care and Use Committee of Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences. We used adult C57BL/6 mice (Hunan SJA Laboratory Animal Company, Changsha, China) for all experiments. We housed the mice in the appropriate environment with a 12/12h light/dark cycle; we supplied water and food ad libitum.

    2.3. Viral vector injection

    We selected the stereotactic injection coordinates according to Paxinos and Franklin’s The Mouse Brain in Stereotaxic Coordinates, 4th edition.26 The stereotactic coordinates for the MOp were as follows: Anterior–posterior (AP) axis: +1.54mm; medial–lateral (ML) axis: ±1.60mm; and dorsal–ventral (DV) axis: 1.60mm from the bregma; we performed the stereotactic injection process as previously reported.5 For the intravenous administration, we injected the C57BL/6 mice with rAAV-PHP.eB-EF1α-N2cG-WPRE-polyA at a dose of 4×1011VG/mouse.

    2.4. Slice preparation and imaging

    We prepared and imaged the slices according to previously reported methods.8 We soaked the brains overnight in 4% paraformaldehyde solution. After we completed dehydration with 30% sucrose solution, we sectioned the brains into 40μm slices with a microtome (Thermo Fisher Scientific, Waltham, MA, USA), which we collected in antifreeze fluid and stored at 20C until further use. After thorough washing with PBS, we attached all the brain slices to microscope slides, which we then counterstained with DAPI (1:4000) and sealed with 70% glycerol. For imaging, we used a virtual microscopy slide-scanning system (Olympus, VS 120, Tokyo, Japan).

    3. Results and Discussion

    3.1. Establishment of novel retrograde trans-multi-synaptic tracing methods

    We previously reported that brain-wide TVA receptor compensation allows rabies virus to retrograde target projection neurons of a specific brain region.5 Based on similar principles, we used AAV-PHP.eB to achieve brain-wide compensation of RVG, and designed a RV-ΔG-based retrograde trans-multi-synaptic system. The functional elements carried in the AAVs and RV-ΔG are illustrated in Figs. 1(a) and 1(b), respectively. Based on our previous findings,24 we selected the CVS-N2c strain of rabies virus for this study. For retrograde trans-multi-synaptic labeling, we used an AAV virus for brain-wide N2cG compensation and a B19G-pseudotyped CVS-N2c-ΔG (B19G-CVS-N2c-ΔG) (Fig. 1(c)). For retrograde trans-multi-synaptic labeling from cell-type-specific starter cells, we used an AAV virus for brain-wide N2cG compensation, an AAV virus carrying TVA controlled by a cell-type-specific promoter, and an EnvA-pseudotyped CVS-N2c-ΔG (EnvA-CVS-N2c-ΔG) (Fig. 1(d)).

    Fig. 1.

    Fig. 1. Schematic diagram of viral vectors and design of retrograde trans-multi-synaptic tracing. (a) AAVs used for retrograde trans-multi-synaptic tracing: rAAV-PHP.eB for carrying N2cG controlled by EF1α promoter, and rAAV9 for carrying TVA controlled by CaMKIIα promoter. (b) RVs used for retrograde trans-multi-synaptic tracing: B19G-CVS-N2c-ΔG-EGFP for retrograde trans-multi-synaptic tracing, and EnvA-CVS-N2c-ΔG-tdTomato for retrograde trans-multi-synaptic tracing from genetically defined starter neurons. EnvA-CVS-N2c-ΔG-tdTomato could specifically recognize and infect TVA-positive neurons. (c) Schematic diagram of retrograde trans-multi-synaptic tracing strategy. i.v., intravenous injection. (d) Schematic diagram of genetically defined retrograde trans-multi-synaptic tracing strategy. i.v., intravenous injection.

    3.2. New method enables efficient trans-monosynaptic labeling

    To verify whether the helper virus could effectively mediate the trans-synaptic tracing of RV-ΔG, we first used the rAAV-PHP.eB vector for in situ delivery of the N2cG gene. We injected rAAV-PHP.eB-EF1α-N2cG into the primary motor area (MOp); after three weeks, we injected B19G-CVS-N2c-ΔG-EGFP at the same site (Fig. 2(a)). Five days later, we sacrificed the mice, and we imaged the upstream inputs of the MOp. We observed many EGFP-positive neurons in the brain regions, such as in the ventral anterior–lateral complex of the thalamus (VAL), the superior colliculus, motor-related (SCm), the substantia nigra, reticular (SNr), and the dentate nucleus (DN) (Fig. 2(b)). These results verified the effect of AAV-PHP.eB as a helper virus.

    Fig. 2.

    Fig. 2. Local injection to verify trans-synaptic effect. (a) rAAV-PHP.eB-EF1α-N2cG was injected in primary motor area (MOp) of C57BL/6 mice; at three weeks postinjection, B19G-CVS-N2c-ΔG-EGFP was injected into the same site, and mice were sacrificed after five days. (b) Fluorescence expression in brain regions upstream of MOp. VAL, ventral anterior–lateral complex of the thalamus; SCm, superior colliculus, motor-related; SNr, substantia nigra, reticular; DN, dentate nucleus. Scale bar =100μm.

    3.3. CVS-N2c-ΔG-based retrograde trans-multi-synaptic tracing

    To trace the upstream input networks of a special brain region, N2cG was carried by rAAV-PHP.eB and compensated in the whole brain via intravenous administration. Three weeks after expression, we injected B19G-CVS-N2c-ΔG-EGFP in the MOp (Fig. 3(a)). Five days later, we sacrificed the mice, and we imaged the upstream inputs of the MOp. We found that B19G-CVS-N2c-ΔG-EGFP infected the injection site (Fig. 3(b)) and retrograde trans-synaptic label the upstream brain regions (Figs. 3(b) and 3(c)), including the VAL, basolateral amygdalar nucleus (BLA), subiculum (SUB), SNr, simple lobule (SIM), DN, and caudate putamen (CPu), among others. The CPu was proved to be an indirect input of the MOp.27,28 These results indicated that CVS-N2c-ΔG-based retrograde trans-multi-synaptic tracing technology can be used to label brain-wide input networks of specific brain regions. When tracing neural circuits in the motor systems of nonhuman primates, the CVS-N2c strain of rabies virus can rapidly spread across one synaptic step every 10–12h.16,29,30 Thus, CVS-N2c-ΔG also has the potential to attain time-dependent retrograde trans-synaptic labeling. In addition, using CVS-N2c-ΔG vector to express a slow fluorescent timer (sFT), whose fluorescence color can change from blue to red over time,2,31 will allow us to distinguish first-, second-, and third-order neurons according to the red-to-blue fluorescence intensity ratio.

    Fig. 3.

    Fig. 3. CVS-N2c-ΔG-based retrograde trans-multi-synaptic tracing at five days postinfection. (a) Schematic diagram of retrograde trans-multi-synaptic tracing. Helper virus expressing N2cG was intravenously administered; after three weeks, the B19G-CVS-N2c-ΔG-EGFP was injected into the primary motor area (MOp). Then mice were sacrificed after five days. (b) Fluorescence expression at injection site (MOp), and upstream brain region caudate putamen (CPu). In large picture on left, scale bar = 1mm; in small pictures on right, scale bar =200μm. (c) Fluorescence expression in upstream brain regions. VAL, ventral anterior–lateral complex of the thalamus; BLA, basolateral amygdalar nucleus; SUB, subiculum; SNr, substantia nigra; SIM, reticular simple lobule; DN, dentate nucleus. Large picture on left, scale bar =1mm; small pictures on right, scale bar =100μm.

    3.4. CVS-N2c-ΔG-based retrograde trans-multi-synaptic tracing from genetically defined starter neurons

    To further trace the input networks of excitatory neurons in the MOp, we administered rAAV9 carrying TVA and EGFP under the control of the CaMKIIα promoter into the MOp. Concomitantly, we used rAAV-PHP.eB for brain-wide N2cG compensation via tail vein injection. After three weeks, we injected EnvA-CVS-ΔG-tdTomato into the MOp (Fig. 4(a)). Five days later, we sacrificed the mice, and we imaged the upstream inputs of the MOp. We observed TVA-positive cells (rAAV, green signals) and tdTomato-positive cells (RV, red signals) in the MOp (Fig. 4(b)). Many upstream brain regions were also labeled by CVS-ΔG-tdTomato (Figs. 4(b) and 4(c)), such as the anterior cingulate area (ACA), VAL, SCm, SNr, and CPu, among others. These results demonstrated that using EnvA-pseudotyped RV-ΔG in combination with rAAV-PHP.eB carrying TVA controlled by a cell-type-specific promoter, retrograde trans-multi-synaptic tracing from genetically defined starter neurons can be achieved. Based on this principle, researchers can also use a Cre-dependent AAV vector expressing TVA, combined with Cre transgenic animals, to create a strategy for the retrograde trans-multi-synaptic tracing from genetically defined starter neurons. Notably, using our strategies to trace the neural circuits of nonhuman primates, AAVs that are more efficient across the blood–brain barrier in nonhuman primates are required. The AAVs may be AAV-B10,32 AAV.CPP.16,33 etc.

    Fig. 4.

    Fig. 4. CVS-N2c-ΔG-based starting-cells-specific retrograde trans-multi-synaptic tracing. (a) Schematic diagram of starting-cells-specific retrograde trans-multi-synaptic tracing. rAAV-PHP.eB expressing N2cG was intravenously administered, and rAAV9 expressing TVA was injected in primary motor area (MOp); after three weeks, EnvA-CVS-N2c-ΔG-tdTomato was injected at same site, then mice were sacrificed after five days. (b) Green fluorescence expression at injection site (MOp), and tdTomato fluorescence expression at the upstream brain region caudate putamen (CPu). Large picture on left, scale bar = 1mm; small pictures on right, scale bar =200μm. (c) tdTomato fluorescence expression in upstream brain regions. ACA, anterior cingulate area; VAL, ventral anterior–lateral complex of the thalamus; SCm, superior colliculus, motor-related; SNr, substantia nigra. Large picture on left, scale bar =1mm; small pictures on right, scale bar =100μm.

    4. Conclusions

    Using a BBB-crossing rAAV-PHP.eB as a helper virus for brain-wide compensation of RVG, RV-ΔG can be used to retrograde trace neural circuits of specific brain regions across multiple synapses. Furthermore, in combination with TVA controlled by a cell-type-specific promoter, RV-ΔG pseudotyped with EnvA enables effective retrograde trans-multi-synaptic spreading from cell-type-specific starter neurons. This study provides new approaches for identifying the input neural networks of rodents and nonhuman primates.

    Acknowledgments

    This study was supported by the STI2030-Major Projects (Grant No. 2021ZD0201003), the National Natural Science Foundation of China (Grant Nos. 31830035, 31771156, 21921004, and 32100899), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB32030200), and the Shenzhen Key Laboratory of Viral Vectors for Biomedicine (Grant No. ZDSYS20200811142401005), the Key Laboratory of Quality Control Technology for Virus- Based Therapeutics, Guangdong Provincial Medical Products Administration (Grant No. 2022ZDZ13), and Open Project Program of Wuhan National Laboratory for Optoelectronics (Grant No. 2019WNLOKF022).

    We are grateful to Liting Luo and Lingling Xu (Core Facility Center, Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences) for technical assistance with centrifugation and microscope imaging. The schematic diagrams in this paper were created with BioRender.com.

    Conflicts of Interest

    The authors declare that there are no conflicts of interest relevant to this paper.

    References

    • 1. T. R. Insel , “Rethinking schizophrenia,” Nature 468, 187–193 (2010). Crossref, Web of ScienceGoogle Scholar
    • 2. K. Lin, X. Zhong, M. Ying, L. Li, S. Tao, X. Zhu, X. He, F. Xu , “A mutant vesicular stomatitis virus with reduced cytotoxicity and enhanced anterograde trans-synaptic efficiency,” Mol. Brain 13, 45 (2020). Crossref, Web of ScienceGoogle Scholar
    • 3. Z. Han, N. Luo, J. Kou, L. Li, W. Ma, S. Peng, Z. Xu, W. Zhang, Y. Qiu, Y. Wu, J. Wang, C. Ye, K. Lin, F. Xu, “AAV11 permits efficient retrograde targeting of projection neurons,” bioRxiv (2022). Google Scholar
    • 4. A. R. Weiss, W. A. Liguore, J. S. Domire, D. Button, J. L. McBride , “Intra-striatal AAV2.retro administration leads to extensive retrograde transport in the rhesus macaque brain: Implications for disease modeling and therapeutic development,” Sci. Rep. 10, 6970 (2020). Crossref, Web of ScienceGoogle Scholar
    • 5. Z. Han, N. Luo, J. Kou, L. Li, Z. Xu, S. Wei, Y. Wu, J. Wang, C. Ye, K. Lin, F. Xu , “Brain-wide TVA compensation allows rabies virus to retrograde target cell-type-specific projection neurons,” Mol. Brain 15, 13 (2022). Crossref, Web of ScienceGoogle Scholar
    • 6. D. G. Tervo, B. Y. Hwang, S. Viswanathan, T. Gaj, M. Lavzin, K. D. Ritola, S. Lindo, S. Michael, E. Kuleshova, D. Ojala, C. C. Huang, C. R. Gerfen, J. Schiller, J. T. Dudman, A. W. Hantman, L. L. Looger, D. V. Schaffer, A. Y. Karpova , “A designer AAV variant permits efficient retrograde access to projection neurons,” Neuron 92, 372–382 (2016). Crossref, Web of ScienceGoogle Scholar
    • 7. J. Tordo, C. O’Leary, A. Antunes, N. Palomar, P. Aldrin-Kirk, M. Basche, A. Bennett, Z. D’Souza, H. Gleitz, A. Godwin, R. J. Holley, H. Parker, A. Y. Liao, P. Rouse, A. S. Youshani, L. Dridi, C. Martins, T. Levade, K. B. Stacey, D. M. Davis, A. Dyer, N. Clement, T. Bjorklund, R. R. Ali, M. Agbandje-McKenna, A. A. Rahim, A. Pshezhetsky, S. N. Waddington, R. M. Linden, B. W. Bigger, E. Henckaerts , “A novel adeno-associated virus capsid with enhanced neurotropism corrects a lysosomal transmembrane enzyme deficiency,” Brain J. Neurol. 141, 2014–2031 (2018). Crossref, Web of ScienceGoogle Scholar
    • 8. K. Lin, X. Zhong, L. Li, M. Ying, T. Yang, Z. Zhang, X. He, F. Xu , “AAV9-Retro mediates efficient transduction with axon terminal absorption and blood-brain barrier transportation,” Mol. Brain 13, 138 (2020). Crossref, Web of ScienceGoogle Scholar
    • 9. S. Tanabe, K. I. Inoue, H. Tsuge, S. Uezono, K. Nagaya, M. Fujiwara, S. Kato, K. Kobayashi, M. Takada , “The use of an optimized chimeric envelope glycoprotein enhances the efficiency of retrograde gene transfer of a pseudotyped lentiviral vector in the primate brain,” Neurosci. Res. 120, 45–52 (2017). Crossref, Web of ScienceGoogle Scholar
    • 10. S. J. Li, A. Vaughan, J. F. Sturgill, A. Kepecs , “A viral receptor complementation strategy to overcome CAV-2 tropism for efficient retrograde targeting of neurons,” Neuron 98, 905–917 e5 (2018). Crossref, Web of ScienceGoogle Scholar
    • 11. E. M. Callaway, L. Luo , “Monosynaptic circuit tracing with glycoprotein-deleted rabies viruses,” J. Neurosci. 35, 8979–8985 (2015). Crossref, Web of ScienceGoogle Scholar
    • 12. P. Su, M. Ying, Z. Han, J. Xia, S. Jin, Y. Li, H. Wang, F. Xu , “High-brightness anterograde transneuronal HSV1 H129 tracer modified using a Trojan horse-like strategy,” Mol. Brain 13, 5 (2020). Crossref, Web of ScienceGoogle Scholar
    • 13. G. E. Pickard, C. A. Smeraski, C. C. Tomlinson, B. W. Banfield, J. Kaufman, C. L. Wilcox, L. W. Enquist, P. J. Sollars , “Intravitreal injection of the attenuated pseudorabies virus PRV Bartha results in infection of the hamster suprachiasmatic nucleus only by retrograde transsynaptic transport via autonomic circuits,” J. Neurosci. 22, 2701–2710 (2002). Crossref, Web of ScienceGoogle Scholar
    • 14. L. A. Schwarz, K. Miyamichi, X. J. Gao, K. T. Beier, B. Weissbourd, K. E. DeLoach, J. Ren, S. Ibanes, R. C. Malenka, E. J. Kremer, L. Luo , “Viral-genetic tracing of the input-output organization of a central noradrenaline circuit,” Nature 524, 88–92 (2015). Crossref, Web of ScienceGoogle Scholar
    • 15. Q. Liu, Y. Wu, H. Wang, F. Jia, F. Xu , “Viral tools for neural circuit tracing,” Neurosci. Bull. 38(12), 1508–1518 (2022). Crossref, Web of ScienceGoogle Scholar
    • 16. E. M. Callaway , “Transneuronal circuit tracing with neurotropic viruses,” Curr. Opin. Neurobiol. 18, 617–623 (2008). Crossref, Web of ScienceGoogle Scholar
    • 17. S. Ohara, K. Inoue, M. P. Witter, T. Iijima , “Untangling neural networks with dual retrograde transsynaptic viral infection,” Front. Neurosci. 3, 344–349 (2009). Crossref, Web of ScienceGoogle Scholar
    • 18. F. Jia, P. Lv, H. Miao, X. W. Shi, H. J. Mei, L. Li, X. Q. Xu, S. J. Tao, F. Q. Xu , “Optimization of the fluorescent protein expression level based on pseudorabies virus bartha strain for neural circuit tracing,” Front. Neuroanat. 13, 63 (2019). Crossref, Web of ScienceGoogle Scholar
    • 19. J. DeFalco, M. Tomishima, H. Y. Liu, C. Zhao, X. L. Cai, J. D. Marth, L. Enquist, J. M. Friedman , “Virus-assisted mapping of neural inputs to a feeding center in the hypothalamus,” Science 291, 2608–2613 (2001). Crossref, Web of ScienceGoogle Scholar
    • 20. L. Sun, Y. Tang, K. Yan, J. Yu, Y. Zou, W. Xu, K. Xiao, Z. Zhang, W. Li, B. Wu, Z. Hu, K. Chen, Z. F. Fu, J. Dai, G. Cao , “Differences in neurotropism and neurotoxicity among retrograde viral tracers,” Mol. Neurodegener. 14, 8 (2019). Crossref, Web of ScienceGoogle Scholar
    • 21. I. R. Wickersham, D. C. Lyon, R. J. Barnard, T. Mori, S. Finke, K. K. Conzelmann, J. A. Young, E. M. Callaway , “Monosynaptic restriction of transsynaptic tracing from single, genetically targeted neurons,” Neuron 53, 639–647 (2007). Crossref, Web of ScienceGoogle Scholar
    • 22. I. R. Wickersham, S. Finke, K. K. Conzelmann, E. M. Callaway , “Retrograde neuronal tracing with a deletion-mutant rabies virus,” Nat. Meth. 4, 47–49 (2007). Crossref, Web of ScienceGoogle Scholar
    • 23. T. R. Reardon, A. J. Murray, G. F. Turi, C. Wirblich, K. R. Croce, M. J. Schnell, T. M. Jessell, A. Losonczy , “Rabies virus CVS-N2c(DeltaG) strain enhances retrograde synaptic transfer and neuronal viability,” Neuron 89, 711–724 (2016). Crossref, Web of ScienceGoogle Scholar
    • 24. K. Lin, L. Li, W. Ma, X. Yang, Z. Han, N. Luo, F. Xu, “A rabies virus-based toolkit for efficient retrograde labeling and monosynaptic tracing,” Neural. Regen. Res. 18, 1827–1833 (2023). Google Scholar
    • 25. Y. H. Chen, M. S. Keiser, B. L. Davidson , “Adeno-associated virus production, purification, and titering,” Curr. Protoc. Mouse Biol. 8, e56 (2018). CrossrefGoogle Scholar
    • 26. G. Paxinos, K. Franklin , Paxinos and Franklin’s the Mouse Brain in Stereotaxic Coordinates, Elsevier/Academic Press, Amsterdam (2012). Google Scholar
    • 27. K. T. Beier, A. Saunders, I. A. Oldenburg, K. Miyamichi, N. Akhtar, L. Luo, S. P. Whelan, B. Sabatini, C. L. Cepko , “Anterograde or retrograde transsynaptic labeling of CNS neurons with vesicular stomatitis virus vectors,” Proc. Natl. Acad. Sci. USA 108, 15414–15419 (2011). Crossref, Web of ScienceGoogle Scholar
    • 28. K. T. Beier, A. B. Saunders, I. A. Oldenburg, B. L. Sabatini, C. L. Cepko , “Vesicular stomatitis virus with the rabies virus glycoprotein directs retrograde transsynaptic transport among neurons in vivo,” Front. Neural Circuits 7, 11 (2013). Crossref, Web of ScienceGoogle Scholar
    • 29. R. M. Kelly, P. L. Strick , “Cerebellar loops with motor cortex and prefrontal cortex of a nonhuman primate,” J. Neurosci. 23, 8432–8444 (2003). Crossref, Web of ScienceGoogle Scholar
    • 30. E. Hoshi, L. Tremblay, J. Féger, P. L. Carras, P. L. Strick , “The cerebellum communicates with the basal ganglia,” Nat. Neurosci. 8, 1491–1493 (2005). Crossref, Web of ScienceGoogle Scholar
    • 31. S. Ohara, Y. Sota, S. Sato, K. I. Tsutsui, T. Iijima , “Increased transgene expression level of rabies virus vector for transsynaptic tracing,” PLoS One 12, e0180960 (2017). Crossref, Web of ScienceGoogle Scholar
    • 32. D. Goertsen, N. C. Flytzanis, N. Goeden, M. R. Chuapoco, A. Cummins, Y. Chen, Y. Fan, Q. Zhang, J. Sharma, Y. Duan, L. Wang, G. Feng, Y. Chen, N. Y. Ip, J. Pickel, V. Gradinaru , “AAV capsid variants with brain-wide transgene expression and decreased liver targeting after intravenous delivery in mouse and marmoset,” Nat. Neurosci. 25, 106–115 (2022). Crossref, Web of ScienceGoogle Scholar
    • 33. Y. Yao, J. Wang, Y. Liu, Y. Qu, K. Wang, Y. Zhang, Y. Chang, Z. Yang, J. Wan, J. Liu, H. Nakashima, S. E. Lawler, E. A. Chiocca, C. F. Cho, F. Bei , “Variants of the adeno-associated virus serotype 9 with enhanced penetration of the blood–brain barrier in rodents and primates,” Nat. Biomed. Eng. 6, 1257–1271 (2022). Crossref, Web of ScienceGoogle Scholar