Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Dec 1.
Published in final edited form as: Nat Methods. 2009 Jun;6(6):431–434. doi: 10.1038/nmeth.1331

Versatile P(acman) BAC Libraries for Transgenesis Studies in Drosophila melanogaster

Koen J T Venken 1, Joseph W Carlson 2,8, Karen L Schulze 3,8, Hongling Pan 3, Yuchun He 3, Rebecca Spokony 4, Kenneth H Wan 2, Maxim Koriabine 5, Pieter J de Jong 5, Kevin P White 4, Hugo J Bellen 1,3,6,7,*, Roger A Hoskins 2
PMCID: PMC2784134  NIHMSID: NIHMS112300  PMID: 19465919

Abstract

We constructed Drosophila melanogaster BAC libraries with 21-kb and 83-kb inserts in the P(acman) system. Clones representing 12-fold coverage and encompassing more than 95% of annotated genes were mapped onto the reference genome. These clones can be integrated into predetermined attP sites in the genome using ΦC31 integrase to rescue mutations. They can be modified through recombineering, for example to incorporate protein tags and assess expression patterns.


Genetic model systems such as Drosophila melanogaster are powerful tools for investigating developmental and cell biological processes, properties of inheritance, the molecular underpinnings of behavior, and the molecular bases of disease 1. The approaches used in model systems rely on the identification of mutations in genes and the characterization of the gene products, often aided by transgenesis techniques 2.

We recently developed a new transgenesis platform for D. melanogaster, the P(acman) (P/ΦC31 artificial chromosome for manipulation) system, that allows modification of cloned fragments by recombineering and germ-line transformation of genomic DNA fragments up to 133 kb in length 3. P(acman) combines a conditionally amplifiable BAC 4, the ability to use recombineering in E. coli for retrieval and manipulation of DNA inserts 5, and bacteriophage ΦC31 integrase-mediated germ-line transformation into the D. melanogaster genome 6,7. Clones are maintained at low-copy number to improve plasmid stability and facilitate recombineering, but can be induced to high-copy number for plasmid isolation to facilitate microinjection of embryos. Recombineering can be used to insert protein tags for in vivo protein localization or acute protein inactivation 8, and to create deletions 9 and point mutations 5 for structure/function analysis. ΦC31-mediated transgenesis integrates DNA constructs at specific pre-determined attP sites dispersed throughout the genome 3,6,7,10, eliminating the need to map integration events and reducing variability in expression due to position effects 10. The technique allows rescue of mutations in large genes 3 and facilitates comparative expression analysis of engineered DNA constructs 7,1012. Previously, genomic regions of interest were cloned into P(acman) by gap-repair from available mapped BAC clones 3. Here, we describe a more efficient approach: we constructed two genomic BAC libraries in the P(acman) system and mapped the cloned inserts by alignment of paired end sequences to the reference genome sequence.

We engineered a novel P(acman) BAC vector for construction of genomic libraries (Fig. 1a). In addition to the published features 3, we included a polylinker embedded within a mutant α-lacZ fragment. It became apparent that in the low-copy-number condition necessary to ensure stability of large genomic inserts, standard α-lacZ fragments are expressed at insufficient levels to permit reliable blue-white colony screening. We isolated a mutant with significantly enhanced β-galactosidase activity resulting from a premature stop codon in the α-lacZ fragment (Supplementary Fig. 1) that permits blue-white selection for cloned inserts at low-copy number using an automated colony picking device.

Figure 1. The P(acman) BAC Vector and Mapped Clones in the eve Region.

Figure 1

(a) Map of attB-P(acman)-CmR-BW with partitioning functions (parA parB and parC), low-copy replication functions (repE and oriS), selectable marker (chloramphenicol acetyl transferase, CmR), conditionally inducible origin of replication (oriV) and dominant eye color marker white +. Genomic DNA was cloned into the BamHI site (Supplementary Fig. 2). (b) A 100-kb region surrounding the eve gene (yellow) on chromosome arm 2R. Mapped CHORI-321 and CHORI-322 clones are indicated below the FlyBase R5.9 gene annotation. CH322-103K22, selected for transformation (Supplementary Table 2) and protein tagging (Fig. 2), is indicated in red.

To create a resource for manipulation and analysis of D. melanogaster genes, we constructed two P(acman) libraries with different insert sizes (Supplementary Fig. 2). For analysis of most genes, we used the library with an insert size of 20 kb. Ninety percent of protein-coding gene annotations in D. melanogaster are less than 12.1 kb in length, and a 20 kb insert size should provide sufficient flanking genomic sequence to contain most genes, including regulatory sequences required for normal expression. For analysis of large genes and gene complexes, we constructed a library with an insert size of 80 kb. High molecular weight genomic DNA was prepared from the D. melanogaster strain used to produce the reference genome sequence. The DNA was fragmented by partial restriction digestion, and size fractions in the 20 kb and 80 kb ranges were recovered and cloned separately to produce two genomic BAC libraries. The libraries produced from the 20 kb and 80 kb fractions were designated CHORI-322 and CHORI-321, respectively. We stocked 73,728 CHORI-322 clones and 36,864 CHORI-321 clones.

To map P(acman) BACs on the genome, paired end sequences were determined and aligned to the reference genome sequence. We mapped consistent paired ends of 33,314 CHORI-322 clones representing 4.3-fold coverage of the X chromosome and 5.9-fold coverage of the autosomes, and 12,328 CHORI-321 clones representing 8.2-fold coverage of the X chromosome and 9.3-fold coverage of the autosomes. The mapped paired end sequences show that the average insert sizes of the CHORI-322 and CHORI-321 libraries are 21.0 kb (+/− 4.0 kb) and 83.3 kb (+/− 21.5 kb), respectively. An additional 18,767 CHORI-322 clones and 11,571 CHORI-321 clones were partially mapped to the genome sequence by alignment of one end sequence only. The two libraries together represent deep coverage of the genome and span most annotated genes (Supplementary Table 1). The mapped CHORI-322 and CHORI-321 clones span 88.9% and 99.3% of annotated genes, respectively. P(acman) clones containing genes and genomic regions of interest can be identified through a web-accessible genome browser (http://pacmanfly.org/) (Fig. 1b) and are available for distribution from the BACPAC Resources Center (http://bacpac.chori.org/).

We tested the P(acman) library resource for transformation efficiency using clones encompassing several genes. For each gene, we identified a clone containing substantial flanking sequences biased toward the 5’ end of the gene annotation. These clones are likely to include the regulatory sequences necessary for normal expression of the gene. For small genes (≤12 kb), a CHORI-322 clone was preferred over a CHORI-321 clone, as smaller clones tend to have higher transformation efficiencies 3. When a mapped CHORI-322 clone was not available for a small gene (e.g. hh, vas and shi) or sufficient 5’ regulatory sequence did not appear to be present in a mapped CHORI-322 clone (e.g. jar, lt and cta), we chose a CHORI-321 clone instead. In total, we selected 38 clones from the CHORI-322 library (Table 1) and 24 clones from the CHORI-321 library (Table 2). The largest clone, encompassing Hnf4, has an insert size of 105 kb. Each clone was isolated and tested for integration into a genomic attP docking site, either VK37 on chromosome arm 2L or VK33 on chromosome arm 3L 3, using ΦC31 integrase 6,7. The transformation efficiency of each clone was defined as the percentage of G0 fertile crosses that yielded at least one transgenic animal. We were successful in obtaining at least one transformant for all CHORI-322 clones (Table 1) and 13 of the 24 CHORI-321 clones (Table 2). In addition, 16 of 17 CHORI-322 clones used for recombineering-mediated tagging (see below) were successfully integrated (Supplementary Table 2). Moreover, 53 of 72 CHORI-321 clones have been integrated successfully in an independent experiment to generate a set of duplication lines, each carrying a clone from a tiling path of overlapping CHORI-321 clones spanning the entire X chromosome (Ellen Popodi and Thom Kaufman, personal communication). These data show that more than 98% (54/55) of CHORI-322 clones and at least 68% (66/96) of CHORI-321 clones can be successfully integrated. For all transformants, the presence of the expected DNA fragment sizes at the integration junctions - indicative of site-specific integration at the respective docking site - was confirmed by multiplex PCR that tests simultaneously for the presence of attP, attB, attR and attL sites (Supplementary Fig. 3).

Table 1. Characterization of CHORI-322 Clones.

Genes contained in 38 CHORI-322 clones are indicated. For each clone, the deduced genomic insert length in bp (insert), attP VK docking site used (VK#), number of fertile G0 crosses (G0), number of vials resulting in at least one transgenic animal (Tr) and integration efficiency (%) are indicated. Note that only the gene of interest is shown; most clones contain more than one gene.

Gene(s) Clone Insert VK# G0 Tr %
Act42A 12M14 19,321 33 84 2 2.4%
Act87E 158M20 19,317 37 65 5 7.7%
alphaTub84B 158D05 20,091 37 51 1 2.0%
bcd 100D18 18,452 37 59 4 6.8%
CG14438 191E24 20,115 33 103 4 3.9%
CG6017 55J22 19,815 37 24 3 12.5%
Chc 123J21 17,647 33 57 8 14.0%
Clc 92D22 20,882 37 32 2 6.3%
Csp
Csp
Csp
06D09
06D09
06D09
23,790
23,790
23,790
16 44 5 11.4%
22 11 2 18.2%
37 115 0 0.0%
Dap160 154I22 22,027 33 73 1 1.4%
DIP1 146O15 20,342 33 69 2 2.9%
Drp1 83H15 19,373 33 115 2 1.7%
endoA 19L12 20,365 37 25 1 4.0%
Eps-15 150F15 21,222 33 61 1 1.6%
ERR 54A09 20,644 37 70 1 1.4%
His2Av 97I07 20,846 37 51 1 2.0%
Hr4 137G06 22,025 33 56 3 5.4%
Hr96 155C21 22,752 37 38 2 5.3%
Khc 162G07 21,662 33 63 1 1.6%
ncd 118A01 22,196 37 104 4 3.8%
Nmnat 175G15 24,376 37 87 2 2.3%
Nrx-IV 154P15 21,688 37 45 3 6.7%
n-syb 83G13 21,536 37 48 1 2.1%
ogre 155A19 22,329 33 109 1 0.9%
Pak 05M18 21,251 37 76 4 5.3%
pen 08M17 22,049 33 71 1 1.4%
piwi 103C03 24,728 33 43 2 4.7%
polo 104P12 23,954 37 28 1 3.6%
Rab5 97N16 20,416 33 59 4 6.8%
sec15/Rab11 152E24 21,979 37 58 2 3.4%
sens 01N16 19,175 37 49 9 18.4%
spn-E 93D04 23,226 37 39 2 5.1%
sqh 130G10 20,074 33 65 8 12.3%
stau 15P05 20,978 33 57 3 5.3%
synj 188H18 18,311 33 36 1 2.8%
Syx1A 142K16 18,384 37 45 1 2.2%
Vha100-1 119J05 19,299 37 53 3 5.7%
wg
wg
wg
192I14
192I14
192I14
22,791
22,791
22,791
13 21 0 0.0%
31 35 5 14.3%
33 78 0 0.0%

Table 2. Characterization of CHORI-321 Clones.

Genes contained in 24 CHORI-321 clones are indicated Colum headings are identical to those in Table 1.

Gene(s) Clone Insert VK# G0 Tr %
cac 60D21 77,150 33 62 2 3.2%
cta 03L03 83,080 33 65 1 1.5%
cta 04I17 91,022 33 81 1 1.2%
dpp 23O18 86,898 33 41 1 2.4%
Dscam 22M14 87,314 33 99 2 2.0%
eag 77E01 97,072 33 56 1 1.8%
ftz-f1 47I12 92,406 37 51 0 0.0%
gfa 57O14 77,006 33 76 0 0.0%
hh 61H05 101,201 37 61 0 0.0%
Hnf4 12P12 104,925 33 63 1 1.6%
Hr38 25N09 85,274 33 107 0 0.0%
Hr46 23L02 86,475 33 62 1 1.6%
jar 76B03 92,518 37 46 0 0.0%
lt 16H04 92,084 33 66 6 9.1%
lt 64G01 92,368 33 60 4 6.7%
lt/cta 05E14 78,102 33 64 2 3.1%
para 18K02 98,254 33 68 0 0.0%
rl 36L01 53,704 33 63 0 0.0%
rl 81D16 102,491 33 79 0 0.0%
shakB 27E22 88,712 33 76 2 2.6%
shi 71G22 70,514 33 53 0 0.0%
syt 08F02 82,244 33 59 0 0.0%
tweek 79N05 76,842 33 69 8 11.6%
vas 69O09 80,238 33 62 0 0.0%

The range of integration efficiencies observed is surprisingly broad. Efficiencies ranged from 0% to 28.1 % for CHORI-322 clones and from 0% to 11.6 % for CHORI-321 clones. The insert sizes of CHORI-322 clones are very similar to each other, so the observed range suggests that some fragments are less efficiently transformed than others due to sequence content or specific interference between certain fragments and docking sites (e.g. Csp and wg). Notably, the high efficiency observed for some CHORI-321 clones (e.g. CH321-16H04, CH321-64G01 and CH321-79N05) suggests that further optimization of the integration efficiency of large clones is possible.

We tested transgenic insertions of ten CHORI-322 and six CHORI-321 clones for their ability to complement lethal mutations in genes. All CHORI-322 clones tested, encompassing the genes CG6017, chc, dap160, drp1, endo, Eps15, n-syb, sqh, synj and vha100-1, rescue lethal mutations in the corresponding genes. To our knowledge, rescue of mutations in endo, n-syb and vha100-1 using genomic fragments has not been reported previously. Similarly, CHORI-321 clones encompassing the genes cac, Dscam, lt and shakB complement lethal mutations in the corresponding genes. Rescue of cac, lt and shakB using genomic fragments has also not been reported previously. Rescue of a lethal mutation in lt with a 92 kb genomic fragment inserted in euchromatin is surprising, because full expression of lt and several other heterochromatic genes has been shown to be dependent on their heterochromatic context 13. Only one of three clones tested complemented lt lethality, suggesting that essential regulatory elements or sufficient genomic context were absent in the other two clones.

To test the utility of recombineering in P(acman) BACs, we introduced EGFP reporter tags into 17 genes encoding transcription factors with well-documented embryonic expression patterns. We inserted the coding region of EGFP in-frame at the 3’ end of the open reading frame, replacing the stop codon and creating C-terminal protein fusions 14 (Supplementary Fig. 4). Both the untagged and tagged constructs were tested for integration using ΦC31 integrase (Supplementary Table 2). Eleven tagged constructs were tested for expression of the fusion protein. Since this EGFP does not fold efficiently in embryos prior to stage 15, we performed immunohistochemistry on embryos with an anti-GFP antibody (Fig. 2 and Supplementary Fig. 5a,b). EGFP fluorescence could be used to visualize fusion protein expression in live embryos only in the late stages of embryonic development (Supplementary Fig. 5c). The expression patterns of eve, D, cad, Dfd, tll, slp2 , and exd are reproduced by the transgenic fusion constructs (Supplementary Discussion). The en and h gene expression patterns appeared to be exceptions (Fig. 2k–l). For h, only two stripes (1 and 5) of expression in the embryo were observed, instead of eight 15. Interestingly, enhancers for stripes 1 and 5 are located in the 7 kb region proximal to the transcription start site, whereas the regulatory elements for the other stripes are located more distally 16. The latter regulatory elements are lacking in CH322-135D17 used to tag h. Hence, the tagged construct is expressed in the expected pattern. Similarly, en expression was only observed in 13 stripes and not the head region 17. This may be due to the absence of regulatory regions in the en clone CH322-92I14 (Judith Kassis, personal communication). These experiments show that recombineering-mediated deletion of genomic sequences in P(acman) constructs can be used to dissect the control of transcription by cis-regulatory elements.

Figure 2. Expression of EGFP Fusion Proteins in Transgenic Embryos.

Figure 2

Fusion proteins were detected using an anti-GFP antibody and peroxidase staining. (a-d) Expression of an even skipped fusion construct recapitulates the native pattern of Eve expression. (a) embryonic stage 5, (b) embryonic stage 9, (c) embryonic stage 11, (d) embryonic stage 15. (e) Dichaete, embryonic stage 5. (f) caudal, embryonic stage 9. (g) Deformed, embryonic stage 11. (h) tailless, embryonic stage 5. (i) sloppy paired 2, embryonic stage 17. (j) extradenticle, embryonic stage 15. (k) engrailed, embryonic stage 9. (l) hairy, embryonic stage 6. Scale bar (a) indicates 50 µm.

In conclusion, we have described a versatile P(acman) BAC library resource for functional analysis of transgenes in D. melanogaster(Supplementary Discussion). We conservatively estimate that the new resource enables in vivo analysis of more than 95% of D. melanogaster genes including large genes, gene complexes and heterochromatic genes (Supplementary Fig. 6). Moreover, protein tagging should prove a valuable alternative to antibody production, particularly when proteins are poorly immunogenic. Finally, the flexibility of recombineering 5 permits the integration of a variety of protein tags for numerous applications 18. The few genes and gene complexes that are too large to be contained within clones in the P(acman) libraries or are otherwise not represented in them can be obtained using the previously described gap-repair procedure 3 and previously mapped and end-sequenced BAC libraries constructed from the same isogenized strain 19,20.

Supplementary Material

1
2

ACKNOWLEDGMENTS

We thank the Washington University Genome Sequencing Center for their excellent BAC end sequencing services. We thank the Bloomington Drosophila Stock Center, NCI-Frederick, N. Copeland (NCI Frederick), A. Hyman (Max Planck Institute), R. Karess (CNRS), R. Ordway (Penn State University), J. Reinitz (Stony Brook University), D. Schmucker (Harvard Medical School), T. Schwarz (Children’s Hospital, Boston), B. Wakimoto (University of Washington), S. Warming (NCI Frederick) and L. Zipursky (UCLA) for reagents. We are especially thankful to J. Bischof, K. Basler (University of Zurich) and F. Karch (University of Geneva) for providing germ-line ΦC31 sources and information about their use. We thank J. Cohen for help with recombineering, N. Giagtzoglou and A. Rajan for help with microscopy, C. Amemiya and D. Frisch for helpful communications and discussions. We are grateful to B. Wakimoto for critical reading of the manuscript. Confocal microscopy was supported by the BCM Intellectual and Developmental Disabilities Research Center. This work was supported by a grant from the Howard Hughes Medical Institute to H.J.B. and the NIH modENCODE project in collaboration with K.P.W. H.J.B. is an Investigator of the Howard Hughes Medical Institute.

Footnotes

ACCESSION NUMBERS

The sequence of the attB-P(acman)-CmR-BW vector and the P(acman) BAC end sequences have been deposited in GenBank under accession numbers FJ931533 and FI329972 to FI494724, respectively.

METHODS

Methods and associated references are available as supplementary online material at http://www.nature.com/naturemethods/.

NOTE

Supplementary information is available on the Nature Methods website.

REFERENCES

  • 1.Bier E. Drosophila, the golden bug, emerges as a tool for human genetics. Nat. Rev. Genet. 2005;6:9–23. doi: 10.1038/nrg1503. [DOI] [PubMed] [Google Scholar]
  • 2.Venken KJ, Bellen HJ. Transgenesis upgrades for Drosophila melanogaster. Development. 2007;134:3571–3584. doi: 10.1242/dev.005686. [DOI] [PubMed] [Google Scholar]
  • 3.Venken KJ, He Y, Hoskins RA, Bellen HJ. P[acman]: a BAC transgenic platform for targeted insertion of large DNA fragments in D. melanogaster. Science. 2006;314:1747–1751. doi: 10.1126/science.1134426. [DOI] [PubMed] [Google Scholar]
  • 4.Wild J, Hradecna Z, Szybalski W. Conditionally amplifiable BACs: switching from single-copy to high-copy vectors and genomic clones. Genome Res. 2002;12:1434–1444. doi: 10.1101/gr.130502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Sawitzke JA, et al. Recombineering: in vivo genetic engineering in E. coli, S. enterica, and beyond. Methods Enzymol. 2007;421:171–199. doi: 10.1016/S0076-6879(06)21015-2. [DOI] [PubMed] [Google Scholar]
  • 6.Groth AC, Fish M, Nusse R, Calos MP. Construction of transgenic Drosophila by using the site-specific integrase from phage ΦC31. Genetics. 2004;166:1775–1782. doi: 10.1534/genetics.166.4.1775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Bischof J, Maeda RK, Hediger M, Karch F, Basler K. An optimized transgenesis system for Drosophila using germ-line-specific ΦC31 integrases. Proc. Natl. Acad. Sci. U.S.A. 2007;104:3312–3317. doi: 10.1073/pnas.0611511104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Venken KJ, et al. Recombineering-mediated tagging of Drosophila genomic constructs for in vivo localization and acute protein inactivation. Nucleic Acids Res. 2008;36:e114. doi: 10.1093/nar/gkn486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Pepple KL, et al. Two-step selection of a single R8 photoreceptor: a bistable loop between senseless and rough locks in R8 fate. Development. 2008;135:4071–4079. doi: 10.1242/dev.028951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Markstein M, Pitsouli C, Villalta C, Celniker SE, Perrimon N. Exploiting position effects and the gypsy retrovirus insulator to engineer precisely expressed transgenes. Nat Genet. 2008;40:476–483. doi: 10.1038/ng.101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Ni JQ, et al. Vector and parameters for targeted transgenic RNA interference in Drosophila melanogaster. Nat Methods. 2008;5:49–51. doi: 10.1038/nmeth1146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Pfeiffer BD, et al. Tools for neuroanatomy and neurogenetics in Drosophila. Proc. Natl. Acad. Sci. U. S.A. 2008;105:9715–9720. doi: 10.1073/pnas.0803697105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Yasuhara JC, Wakimoto BT. Oxymoron no more: the expanding world of heterochromatic genes. Trends Genet. 2006;22:330–338. doi: 10.1016/j.tig.2006.04.008. [DOI] [PubMed] [Google Scholar]
  • 14.Poser I, et al. BAC TransgeneOmics: a high-throughput method for exploration of protein function in mammals. Nat Methods. 2008;5:409–415. doi: 10.1038/nmeth.1199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Hooper KL, Parkhurst SM, Ish-Horowicz D. Spatial control of hairy protein expression during embryogenesis. Development. 1989;107:489–504. doi: 10.1242/dev.107.3.489. [DOI] [PubMed] [Google Scholar]
  • 16.Howard KR, Struhl G. Decoding positional information: regulation of the pair-rule gene hairy. Development. 1990;110:1223–1231. doi: 10.1242/dev.110.4.1223. [DOI] [PubMed] [Google Scholar]
  • 17.DiNardo S, Kuner JM, Theis J, O'Farrell PH. Development of embryonic pattern in D. melanogaster as revealed by accumulation of the nuclear engrailed protein. Cell. 1985;43:59–69. doi: 10.1016/0092-8674(85)90012-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Giepmans BN, Adams SR, Ellisman MH, Tsien RY. The fluorescent toolbox for assessing protein location and function. Science. 2006;312:217–224. doi: 10.1126/science.1124618. [DOI] [PubMed] [Google Scholar]
  • 19.Adams MD, et al. The genome sequence of Drosophila melanogaster. Science. 2000;287:2185–2195. doi: 10.1126/science.287.5461.2185. [DOI] [PubMed] [Google Scholar]
  • 20.Hoskins RA, et al. Sequence finishing and mapping of Drosophila melanogaster heterochromatin. Science. 2007;316:1625–1628. doi: 10.1126/science.1139816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Lee EC, et al. A highly efficient Escherichia coli-based chromosome engineering system adapted for recombinogenic targeting and subcloning of BAC DNA. Genomics. 2001;73:56–65. doi: 10.1006/geno.2000.6451. [DOI] [PubMed] [Google Scholar]
  • 22.Hobert O. PCR fusion-based approach to create reporter gene constructs for expression analysis in transgenic C. elegans. Biotechniques. 2002;32:728–730. doi: 10.2144/02324bm01. [DOI] [PubMed] [Google Scholar]
  • 23.Warming S, Costantino N, Court DL, Jenkins NA, Copeland NG. Simple and highly efficient BAC recombineering using galK selection. Nucleic Acids Res. 2005;33:e36. doi: 10.1093/nar/gni035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Kim UJ, et al. Construction and characterization of a human bacterial artificial chromosome library. Genomics. 1996;34:213–218. doi: 10.1006/geno.1996.0268. [DOI] [PubMed] [Google Scholar]
  • 25.Rubin GM, Spradling AC. Genetic transformation of Drosophila with transposable element vectors. Science. 1982;218:348–353. doi: 10.1126/science.6289436. [DOI] [PubMed] [Google Scholar]
  • 26.Osoegawa K, et al. An improved approach for construction of bacterial artificial chromosome libraries. Genomics. 1998;52:1–8. doi: 10.1006/geno.1998.5423. [DOI] [PubMed] [Google Scholar]
  • 27.Brizuela BJ, Elfring L, Ballard J, Tamkun JW, Kennison JA. Genetic analysis of the brahma gene of Drosophila melanogaster and polytene chromosome subdivisions 72AB. Genetics. 1994;137:803–813. doi: 10.1093/genetics/137.3.803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Hoskins RA, et al. A BAC-based physical map of the major autosomes of Drosophila melanogaster. Science. 2000;287:2271–2274. doi: 10.1126/science.287.5461.2271. [DOI] [PubMed] [Google Scholar]
  • 29.Altschul SF, et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25:3389–3402. doi: 10.1093/nar/25.17.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Stein LD, et al. The generic genome browser: a building block for a model organism system database. Genome Res. 2002;12:1599–1610. doi: 10.1101/gr.403602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Gong S, et al. A gene expression atlas of the central nervous system based on bacterial artificial chromosomes. Nature. 2003;425:917–925. doi: 10.1038/nature02033. [DOI] [PubMed] [Google Scholar]
  • 32.Ohyama T, et al. Huntingtin-interacting protein 14, a palmitoyl transferase required for exocytosis and targeting of CSP to synaptic vesicles. J Cell Biol. 2007;179:1481–1496. doi: 10.1083/jcb.200710061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Hiesinger PR, et al. The v-ATPase V0 subunit a1 is required for a late step in synaptic vesicle exocytosis in Drosophila. Cell. 2005;121:607–620. doi: 10.1016/j.cell.2005.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Deitcher DL, et al. Distinct requirements for evoked and spontaneous release of neurotransmitter are revealed by mutations in the Drosophila gene neuronal-synaptobrevin. J Neurosci. 1998;18:2028–2039. doi: 10.1523/JNEUROSCI.18-06-02028.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Verstreken P, et al. Synaptic mitochondria are critical for mobilization of reserve pool vesicles at Drosophila neuromuscular junctions. Neuron. 2005;47:365–378. doi: 10.1016/j.neuron.2005.06.018. [DOI] [PubMed] [Google Scholar]
  • 36.Bazinet C, Katzen AL, Morgan M, Mahowald AP, Lemmon SK. The Drosophila clathrin heavy chain gene: clathrin function is essential in a multicellular organism. Genetics. 1993;134:1119–1134. doi: 10.1093/genetics/134.4.1119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Jordan P, Karess R. Myosin light chain-activating phosphorylation sites are required for oogenesis in Drosophila. J Cell Biol. 1997;139:1805–1819. doi: 10.1083/jcb.139.7.1805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Koh TW, et al. Eps15 and Dap160 control synaptic vesicle membrane retrieval and synapse development. J Cell Biol. 2007;178:309–322. doi: 10.1083/jcb.200701030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Koh TW, Verstreken P, Bellen HJ. Dap160/intersectin acts as a stabilizing scaffold required for synaptic development and vesicle endocytosis. Neuron. 2004;43:193–205. doi: 10.1016/j.neuron.2004.06.029. [DOI] [PubMed] [Google Scholar]
  • 40.Verstreken P, et al. Endophilin mutations block clathrin-mediated endocytosis but not neurotransmitter release. Cell. 2002;109:101–112. doi: 10.1016/s0092-8674(02)00688-8. [DOI] [PubMed] [Google Scholar]
  • 41.Verstreken P, et al. Synaptojanin is recruited by endophilin to promote synaptic vesicle uncoating. Neuron. 2003;40:733–748. doi: 10.1016/s0896-6273(03)00644-5. [DOI] [PubMed] [Google Scholar]
  • 42.Kawasaki F, Collins SC, Ordway RW. Synaptic calcium-channel function in Drosophila: analysis and transformation rescue of temperature-sensitive paralytic and lethal mutations of cacophony. J Neurosci. 2002;22:5856–5864. doi: 10.1523/JNEUROSCI.22-14-05856.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Lindsley DL, Zimm GG. The Genome of Drosophila melanogaster. Academic, San Diego. 1992 [Google Scholar]
  • 44.Kramers PG, Schalet AP, Paradi E, Huiser-Hoogteyling L. High proportion of multi-locus deletions among hycanthone-induced X-linked recessive lethals in Drosophila melanogaster. Mutat Res. 1983;107:187–201. doi: 10.1016/0027-5107(83)90162-8. [DOI] [PubMed] [Google Scholar]
  • 45.Perrimon N, Smouse D, Miklos GL. Developmental genetics of loci at the base of the X chromosome of Drosophila melanogaster. Genetics. 1989;121:313–331. doi: 10.1093/genetics/121.2.313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Hummel T, et al. Axonal targeting of olfactory receptor neurons in Drosophila is controlled by Dscam. Neuron. 2003;37:221–231. doi: 10.1016/s0896-6273(02)01183-2. [DOI] [PubMed] [Google Scholar]
  • 47.Wakimoto BT, Hearn MG. The effects of chromosome rearrangements on the expression of heterochromatic genes in chromosome 2L of Drosophila melanogaster. Genetics. 1990;125:141–154. doi: 10.1093/genetics/125.1.141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Kasprowicz J, et al. Inactivation of clathrin heavy chain inhibits synaptic recycling but allows bulk membrane uptake. J Cell Biol. 2008;182:1007–1016. doi: 10.1083/jcb.200804162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Rasband W. ImageJ: Image processing and analysis in Java. Online. 2009 [Google Scholar]
  • 50.Kosman D, Small S, Reinitz J. Rapid preparation of a panel of polyclonal antibodies to Drosophila segmentation proteins. Dev Genes Evol. 1998;208:290–294. doi: 10.1007/s004270050184. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1
2

RESOURCES