Skip to main content
PMC Canada Author Manuscripts logoLink to PMC Canada Author Manuscripts
. Author manuscript; available in PMC: 2013 Jul 29.
Published in final edited form as: Cytogenet Genome Res. 2007;117(1-4):146–157. doi: 10.1159/000103175

Evolutionary conservation of alternative splicing in chicken

S Katyal 1, Z Gao 1, R-Z Liu 1, R Godbout 1
PMCID: PMC3726401  CAMSID: CAMS3218  PMID: 17675855

Abstract

Alternative splicing represents a source of great diversity for regulating protein expression and function. It has been estimated that one-third to two-thirds of mammalian genes are alternatively spliced. With the sequencing of the chicken genome and analysis of transcripts expressed in chicken tissues, we are now in a position to address evolutionary conservation of alternative splicing events in chicken and mammals. Here, we compare chicken and mammalian transcript sequences of 41 alternatively-spliced genes and 50 frequently accessed genes. Our results support a high frequency of splicing events in chicken, similar to that observed in mammals.


Alternative splicing is the mechanism whereby different transcripts are generated from a single gene. Major alterations in function, regulation, cellular and subcellular localization, and/or abundance, may result from cell-, tissue-and developmental stage-specific alternative splicing events. It is estimated that there are ~30,000 genes in the genomes of higher eukaryotes, including chicken. As many as one-third to two-thirds of mammalian genes are believed to be alternatively spliced resulting in the production of proteins in excess of 100,000 (Croft et al., 2000; Modrek and Lee, 2002; Johnson et al., 2003). Alternatively splicing therefore represents a source of great protein diversity and potential functional diversity in complex organisms.

With the sequencing and analysis of genomes from different species, it has become possible to address the evolutionary conservation of alternative splicing events. Because of the wealth of data available from the human and mouse genomes, most of the studies done to date have involved comparison of these two mammalian species. For example, based on ESTs and mRNAs corresponding to the splice junctions of 2,932 introns from 786 genes, Thanaraj et al. (2003) estimated that 61% of alternative and 74% of constitutive splice junctions are conserved between mouse and human. Analysis of 9,434 human and mouse orthologues by Modrek and Lee (2003) led to the conclusion that exons that are only included in alternative splice forms are poorly conserved in mouse and human, suggesting that they are the products of recent exon creation or loss. In agreement with this study, a survey of 10,818 pairs of human and mouse genes representing 104,103 human splice forms revealed that the majority of human alternative splices are either divergent (49%) or novel (44%) when compared to mouse transcript and genome sequences (Kan et al., 2005). In a recent study, Pan et al. (2005) reported that species-specific alternative splicing of conserved exons is relatively common (>11%) in human and mouse, indicating an important role in the evolutionary differences between mammalian species.

Splicing regulatory elements include the 5′ and 3′ splice sites, the adenine-containing branchpoint sequence, and splicing enhancer and suppressor elements (reviewed in Black, 2003). Small nuclear ribonucleoproteins (snRNPs) and related cofactors bind these cis splicing regulatory elements which, in turn, associate with each other to form the spliceosome. Alternative splicing involves the binding of cis splicing suppressor and enhancer elements by trans factors, including heterogeneous nuclear ribonucleoproteins (hnRNPs) and serine-arginine-rich or related proteins. Factors and elements involved in alternative splicing are well-conserved in higher eukaryotes (Boue et al., 2003; Kondrashov and Koonin, 2003; Lareau et al., 2004; Resch et al., 2004).

Comparison of mammals with the phylogenetically distant chicken offers a unique opportunity to address the functional importance of alternatively spliced products and their evolutionary conservation. Here, we carry out a detailed study of 41 relatively well-studied alternatively spliced genes for which information is available in chicken and mammals. Twenty-five of these genes play a role in the retina, while 16 were selected based on reported alternative splicing in chicken. As a more general strategy, we also carry out a survey of the 50 most frequently accessed genes at the GeneCards homesite to assess the frequency of alternative splicing in human versus chicken genes.

Alternative splicing of genes expressed in the retina

Our first strategy was to identify mammalian genes expressed in the retina which have well-documented alternatively spliced forms. We concentrated on genes expressed in the retina as this tissue is remarkably well-conserved both structurally and functionally in chicken and mammals, suggesting the splicing events may be similarly conserved. The retina is composed of six classes of neuronal cells (ganglion, amacrine, interplexiform, bipolar, horizontal, photoreceptor) and one class of glial cells (Müller) derived from multipotent neuroectodermal precursor cells (Turner and Cepko, 1987; Wetts and Fraser, 1988). The orderly appearance of each class of retinal cells, followed by their migration to distinct parts of the retina, culminates in the formation of three nuclear layers (ganglion, inner and outer), separated by plexiform layers.

We used a combination of approaches, including literature survey, and examination of DNA and EST databases to determine which of the alternative splice variants found in mammals were conserved in chicken (Table 1). Coding regions shown to undergo alternative splicing were used to search chicken EST databases using the ‘tblastn’ search criteria. A total of 508,967 ESTs are listed in the NCBI database (http://www.ncbi.nlm.nih.gov/BLAST/) which includes >40,000 ESTs from the University of Delaware database (http://www.chickest.udel.edu/) and 339,314 ESTs from the BBSRC database (http://www.chick.umist.ac.uk/). When possible, we have divided the genes into categories; e.g. homeobox genes (PAX6, SIX3, OTX2, VSX1), photoreceptor genes (ARB1, GRK1, PDE6), cell signaling genes (RELN, DAB1), glutamate receptors and transporters (GLT1, GRM1, GRIN1), oncogenes and tumor suppressors (MAF, WT1).

Table 1.

Alternative splicing of genes expressed in retina

Gene symbol Gene name Isoforms nt accession #/reference Chicken isoforms
AGRN Agrin agrin 0 (B0) NM_198576 NM_205527
agrin 8 (B8) Rupp et al., 1992 Ruegg et al., 1992
agrin 11 (B11) Rupp et al., 1992 Ruegg et al., 1992
agrin 19 (B19) Rupp et al., 1992 U07271
ARB1/ARRB1 arrestin beta 1 ARRB 1a NM_177231 EST
ARRB 1b NM_178220 Not found
DAB1 Disabled-1 DAB1-L NM_021080 AY242123
DAB1-E Unpublished AY242122
FOXN3 Forkhead box protein N3 FOXN3-alpha AM114794 Not found
FOXN3-beta AM114795 EST
GABRB2 gamma-aminobutyric acid A receptor beta 2 GABAAβ2S NM_000813 Harvey et al., 1994
GABAAβ2L NM_021911 Harvey et al., 1994
GABRG2 gamma-aminobutyric acid A receptor gamma 2 GABAAγ2S NM_000816 Glencorse et al., 1992
GABAAγ2L NM_198904 NM_205345
GABBR1 gamma-aminobutyric acid B receptor 1 GABAB1a NM_001470 EST
GABAB1b NM_021903 Not found
GABAB1c NM_021904 Not found
GABAB1d NM_021905 Not found
GABAB1e Billinton et al., 2001 Not found
GLT1/SLC1A2 glutamate transporter 1 GLT1a NM_004171 EST
GLT1b/v Schmitt et al., 2002 NM_001012899
GLT1c Rauen et al., 2004 Not found
GRK1 G-protein receptor kinase 1 (rhodopsin kinase) GRK1a NM_002929 NM_204695
GRK1b Zhao et al., 1998 Not found
MAF musculoaponeurotic fibrosarcoma oncogene c-Maf-alpha NM_005360 D28596
c-Maf-beta NM_001031804 D28596
GRM1 metabotropic glutamate receptor 1 mGluR1a NM_000838 EST
mGluR1b HSU31216/gi:945098 EST
mGluR1c Pin et al., 1992 Not found
mGluR1d Laurie et al., 1996 Not found
mGluR1g Makoff et al., 1997 Not found
NETO1 neuropilin- and tolloid-like protein 1 NETO1 isoform 1 NM_138999 Not found
NETO1 isoform 2 NM_153181 XM_414109
NETO1 isoform 3 NM_138966 XM_419104
NMDAR1/ GR1N1/NR1 NMDA receptor 1/ glutamate receptor, ionotropic, N-methyl D-aspartate 1/Nuclear factor 1 NMDAR1a NM_000832 EST
NMDAR1b NM_021569 NM_206979
NMDAR1c NM_007327 AAY663373
NR2E3 Nuclear receptor subfamily 2, group E, member 3 NR2E3-alpha NM_016346 Not found
NR2E3-beta NM_014249 NM_204594
NRL Neural retina leucine zipper Nrl NM_006177 Not found
Nrl-ins Wistow et al., 2002 Not found
NUMB numb homologue numb isoform 1 NM_001005743 Not found
numb isoform 2 NM_001005744 AF176086
numb isoform 3 NM_003744 AF213691
numb isoform 4 NM_001005745 Not found
OTX2 orthodenticle homolog 2 OTX2A NM_021728 NM_204520
OTX2B NM_172337 EST
PAX6 Paired box gene 6 PAX6-alpha NM_000280 Not found
PAX6-beta NM_001604 NM_205066
PDE6G phosphodiesterase-6 gamma PDE6G NM_002602 NM_204445
PDE6G (–ex2) Wistow et al., 2002 Not found
RPGR retinitis pigmentosa GTPase regulator RPGRa NM_00328 EST
RPGRb NM_001023582 Not found
RPGRc NM_001034853 XM_416780
RELN Reelin RELNa NM_005045 XM_425424
RELNb NM_173054 EST
SIX3 sine oculis homeobox homolog 3 SIX3-alpha Kawakami et al., 1996 Not found
SlX3-beta NM_005413 AY373324
VEGF vascular endothelial growth factor VEGF165a NM_001025368 AY168004
VEGF206a NM_001025366 NM_205042
VEGF165b NM_001033756 Not found
VSX1 visual system homeobox 1 VSX1A NM_014588 NM_204769
VSX1B NM_199425 Not found
WT1 Wilms tumor 1 WT1-beta(−KTS) NM_024424 EST
WT1-delta(+KTS) NM_024426 NM_205216

PAX6, SIX3, OTX2, VSX1

A well-studied gene in the developing retina is the homeobox-containing paired box gene 6 (PAX6). Two alternatively spliced Pax6 isoforms have been identified (Pax6(−5a) and Pax6(+5a)), the longest of which has an exon 5a (14 amino acid) insertion in its paired domain. These two Pax6 isoforms, which are evolutionary conserved in human, mouse and chicken, differ in the DNA binding properties of their respective paired domains (Azuma et al., 2005) (Table 1). Azuma et al. (2005) have shown that ectopic over-expression of Pax6(+5a) in the eyes of developing chick embryos induces a well-differentiated retina-like structure, in keeping with a role for Pax6(+5a) in playing a role specific to a specialized region of the retina called area centralis or fovea where the image of an object is centered.

The sine oculis homeobox homolog 3 (SIX3) is another homeobox gene expressed in the retina. Two alternatively spliced Six3 transcripts have been identified in mouse (Kawakami et al., 1996). Only the longer form of Six3 (Six3-beta) has been reported in humans and chicken. Otx2 (orthodendicle homolog 2) controls photoreceptor cell fate in the developing retina (Nishida et al., 2003). Two Otx2 splice variants have been reported, a long form and a shorter form derived from an alternative 5′ UTR and 3′ splice acceptor site in the second coding exon. Both forms are expressed in chicken based on DNA and EST database searches. A fourth homeobox gene expressed in the retina, VSX1 (visual system homeobox 1), is required for terminal differentiation of cone bipolar cells (Hayashi et al., 2000; Ohtoshi et al., 2004). Of the two Vsx1 splice isoforms reported in human, only one has been identified in chicken.

Arrestin, rhodopsin kinase, PDE6G

In photoreceptors, light activates rhodopsin which in turn activates the G-protein, transducin. Rhodopsin remains active until it becomes phosphorylated by rhodopsin kinase. Arrestin, involved in de-activation of G-protein coupled receptors and cascade inactivation, binds phosphorylated rhodopsin, thus quenching its activity and allowing recovery of the light response. Two alternatively spliced forms of arrestin beta 1 (ARB1) have been identified: p48 and p44. The longer isoform contains 35 amino acids at its C-terminus which are replaced by a single alanine in the truncated form. While both forms of arrestin can bind and quench phosphorylated rhodopsin, the longer form also appears to be able to quench the activity of nonphosphorylated rhodopsin, suggesting a role in dark adaptation (Burns et al., 2006). A search of chicken ESTs resulted in the identification of the long form of arrestin, but not the short form.

Two alternatively-spliced variants of rhodopsin kinase (G-protein receptor kinase 1) have been identified in human: GRK1a and GRK1b (Zhao et al., 1998). The latter includes intron 6 in its coding region. Whereas GRK1a has been described in chicken (Zhao et al., 1999), we were not able to identify cDNAs or ESTs corresponding to GRK1b. Type 6 cGMP-phosphodiesterase (PDE6) is the G-protein-activated effector that regulates levels of cGMP in photoreceptors. The type 6 PDE gamma subunit (PDE6G) has two isoforms generated by alternative splicing, one of which lacks exon 2, resulting in loss of the N-terminal domain required to bind the alpha and beta subunits of PDE (Wistow et al., 2002). We only found evidence for the longer isoform of PDE6G in chicken databases.

DAB1, REELIN

The Reelin-Disabled 1 (Dab1) signaling pathway plays a key role in neuronal cell migration and in the positioning of neurons within laminated structures. Reelin and Dab1 knock-out mice have a number of retinal abnormalities, primarily involving the retinal synaptic circuitry (Rice et al., 2001). We have found that the Disabled-1 (DAB1) gene is alternatively spliced in the developing chick retina (Katyal and Godbout, 2004), resulting in early (E) and late (L) isoforms. Dab1-L differs from Dab1-E by the inclusion of two exons containing two Src family kinase phosphorylation sites. We have evidence that both forms of Dab1 are also present in human retina (Katyal and Godbout, unpublished observations). Additional human and mouse Dab1 splicing variants have been reported in the literature (Bar et al., 2003), none of which are found in chicken databases. Two alternatively spliced transcripts encoding the secreted extracellular matrix glycoprotein Reelin have been reported in humans, both of which appear to be expressed in chicken based on DNA and EST database searches.

GABAA (γ2, β2), GABAB

Gamma aminobutyric acid (GABA) is a major inhibitory neurotransmitter in the retina. Alternative splicing has been observed in a number of protein subunits associated with GABA. For example, two variants of GABAA receptor γ2 subunits (γ2S and γ2L) which differ by eight amino acids encoded by a 24-bp cassette exon, have been described in human, mouse and chicken (Glencorse et al., 1992). Similarly, two alternatively spliced GABAA β2 subunits (β2S and β2L) which differ by 38 amino acids, were identified in human but not in rat or mouse (Harvey et al., 1994; McKinley et al., 1995). Two alternatively spliced variants of GABAA receptor β2 subunit which differ by 17 amino acids rather than 38 amino acids have been reported in chicken (Harvey et al., 1994). There is little similarity in the 17 amino acid insertion found in chicken and the 38 amino acid insertion found in human GABAA receptor β2 subunit. Although five variants of the GABAB receptor subunit 1 have been described in the literature (Isomoto et al., 1998; Schwarz et al., 2000; Benke et al., 2002), we found evidence for only one form (GABAB-1α) expressed in chicken based on database searches.

GLT1, mGluR1, NMDAR1

Three glutamate transporter 1 (GLT) isoforms (GLT1a, GLT1b/v, GLT1c) have been characterized in humans and rats (Schmitt et al., 2002; Rauen et al., 2004). In the retina, GLT1b/v is highly expressed in bipolar cells and cone photoreceptors, whereas GLT1c is primarily found in rod and cone photoreceptors (Rauen et al., 2004). Both the GLT1a and GLT1b/v isoforms are found in chicken based on EST and DNA database searches.

Metabotropic glutamate receptors group 1 (mGluR1) induces slowly arising excitation in postsynaptic neurons and control glutamate release presynaptically. At least four alternatively spliced mGluR1 transcripts have been identified in mammals (Conn and Pin, 1997), two of which (mGluR1a and mGlu1b) were found in chicken ESTs. Transfection of these two variants in chicken retinal cultures demonstrates that targeting of mGluR1 to dendrites and axons is controlled by alternative splicing (Francesconi and Duvoisin, 2002).

Human NMDA receptor 1 (NMDAR1, also called nuclear receptor 1 (NR1) or glutamate receptor, ionotropic, N-methyl D-aspartate 1 (GRIN1)) has three well-characterized alternatively spliced transcripts: NMDAR1a, NMDAR1b, NMDAR1c. Alternative splicing of NMDAR1 involves exons 20 and 21, with NMDAR1a being required for the repair of injured retinal ganglion cells. Chicken NMDAR1b and NMDAR1c cDNA sequences have been entered in the DNA database, and we identified ESTs corresponding to NMDAR1a, suggesting that all three NMDAR1 alternatively spliced products are conserved in mammals and chicken (Kreutz et al., 1998).

Maf, WT1

The musculoaponeurotic fibrosarcoma oncogene Maf has been shown to be expressed in the retina, although its function in this tissue is unknown. Two alternatively spliced forms of Maf (c-Maf-α and c-Maf-β) have been identified in human, mouse and chicken (Huang et al., 2002).

At least four splice variants have been reported for the human Wilms tumor 1 (WT1) gene, involving an alternatively spliced exon 5 and alternative splice donor sites leading to the insertion/deletion of a tripeptide KTS at the end of exon 9 in the C-terminus of the protein (Haber et al., 1991). WT1-beta(−KTS), which lacks the KTS amino acids, is required for retina formation, whereas WT-beta(+KTS) is important for the development of the olfactory epithelium (Wagner et al., 2005). Both the (−KTS) and (+KTS) isoforms of WT1 are found in chicken.

NETO1

Neuropilin- and tolloid-like protein 1 (NETO1) is a retina and brain specific gene encoding a putative transmembrane protein with two extracellular CUB domains. Three alternatively spliced variants are generated through alternate usage of two leader exons (1a and 1b) and exon 5 (Stohr et al., 2002). The soluble isoform 1 which contains only one CUB domain is restricted to the retina. Although isoforms 2 and 3 were identified in the chicken, we found no evidence of isoform 1 expression based on database searches.

Nrl, NR2E3

There are two alternatively spliced Nrl (neural retinal leucine zipper) transcripts in humans: Nrl and Nrl-ins which contains a 59 amino acid alternative exon between exons 2 and 3 (Wistow et al., 2002). To date, there is no evidence of either Nrl isoform in either DNA or EST chicken databases.

Two alternatively spliced forms of nuclear receptor subfamily 2, group E, member 3 (NR2E3) are expressed in different classes of retinal cells: NR2E3-alpha is specifically expressed in photoreceptors, whereas NR2E3-beta is expressed in Müller glial cells as well as in the retinal pigment epithelium (Chen et al., 1999). The chicken NR2E3-beta cDNA sequence has been entered in the DNA database; however, database searches revealed no counterpart to the photoreceptor-specific form in chicken.

ALDH1A1

ALDH1A1 has previously been shown to be specifically expressed in the dorsal retina of chick, mouse and zebrafish during development where it has been postulated to play a role in the generation of a retinoic acid gradient (McCaffery et al., 1991, 1992; Godbout et al., 1996). An ALDH1A1 variant with two additional 5′ untranslated exons, specifically found in the chick retina, gives rise to a retinal ALDH1A1 transcript that is considerably longer than that found in liver (Godbout, 1992; Godbout and Monckton, 2001). We found no evidence of this longer form of ALDH1A1 in either human or mouse.

Agrin

Agrin in a heparin sulfate proteoglycan involved in the formation of acetylcholine receptor aggregates. Agrin isoforms are expressed in the nervous system including the brain, retina, and spinal cord (Kirsch and Kroger, 1996; Kroger, 1997). The agrin transcript is alternatively spliced to include or exclude small exons at three positions. Four agrin splice variants have been identified in rat and chicken: agrin0 (B0), agrin8 (B8), agrin11 (B11) and agrin19 (B19) where the number indicates the number of amino acids inserted as the result of alternative splicing (Ruegg et al., 1992; Rupp et al., 1992; Thomas et al., 1993; O’Connor et al., 1994).

RPGR

Retinitis pigmentosa GTPase regulator (RPGR) is required for the maintenance of photoreceptor viability (Meindl et al., 1996). Mutations in RPGR account for 10–20% of all cases of retinitis pigmentosa (RP) and >70% of X-linked RP. Three main splicing variants of RPGR have been reported in humans: variant a, with multiple alternative exons in the 3′ coding region; variant b, preferentially expressed in the retina, with multiple alternative exons in the 3′ coding region (shortest isoform); and variant c, the longest isoform with an extensive glutamic acid- and glycine-rich domain in its C-terminus. A sequence similar to variant c has been reported in chicken. Furthermore, we have identified ESTs with a low level of similarity to variant a.

VEGF

Multiple splice variants of vascular endothelial growth factor (VEGF) have been identified, some with pro-angiogenic functions, others with anti-angiogenic functions. The defining features of pro-angiogenic VEGF isoforms is inclusion of the six C-terminal amino acids CDKPRR encoded by exon 8, whereas the anti-angiogenic isoforms result from usage of an alternate acceptor site downstream of exon 8 resulting in the six C-terminal amino acids SLTRKD (Bates et al., 2002). Three major pro-angiogenic isoforms have been described: VEGF189, VEGF165 (the dominant form) and VEGF121 (Ferrara and Davis-Smith, 1997). Anti-angiogenic counterparts to these three VEGFs have also been reported: VEGF189b, VEGF165b, VEGF121b (Woolard et al., 2004). Both the pro-angiogenic and anti-angiogenic forms of VEGF are expressed in the retina where they are believed to play a role in diabetic retinopathy (Perrin et al., 2005). Two chicken pro-angiogenic VEGF isoforms VEGF165 and VEGF206 have been annotated in the databases to date. EST database searches failed to reveal anti-angiogenic VEGF isoforms.

NUMB

Numb is a phosphotyrosine-binding (PTB) domain protein involved in cell fate determination in Drosophila. Four mouse Numb isoforms (p72, p71, p66, p65) are generated as the result of alternative splicing. The two longest isoforms, p72 and p71, are expressed in the early retina, while the p66 and p65 isoforms are found in adult retina (Dho et al., 1999; Dooley et al., 2003). Two numb isoforms have been identified in chicken, corresponding to the p66 and p71 isoforms.

Alternative splicing of genes expressed in chicken brain and other tissues

Our second strategy was to examine well-documented alternative splicing events in chicken and to determine whether similar alternative splicing events occur in mammals based on literature searches and sequence comparisons using EST and DNA databases (Table 2). Many of the genes in this category represent genes expressed in brain, likely reflecting a preponderance of alternative splicing events in brain compared to other tissues.

Table 2.

Alternative splicing of genes expressed in chicken brain and other tissues

Gene symbol Gene name Isoforms nt accession # (human)/reference Chicken isoforms
CSNK1A1 casein kinase I CKIalpha NM_001025105 NM_205053
CKIalphaL Yong et al., 2000 Green and Bennett, 1998
CKIalphaS Yong et al., 2000 Green and Bennett, 1998
CKIalphaLS (chicken only) Green and Bennett, 1998
DBN1 drebrin Drebrin E2 NM_004395 NM 205499, Kojima et al., 1993
Drebrin El (chicken only) Toda et al., 1993
DrebrinA NM_080881 Kojima et al., 1993
FGFR2 fibroblast growth factor receptor 2 FGFR2IIIb/K-SAM NM_022970 Sato et al., 1992
FGFR2IIIc/BEK NM_000141 NM_205319, Sato et al., 1991
GRIA4 ionotropic glutamate receptor 4/AMPA4 GluR4a NM_000829 NM_205214
GluR4c Kawahara et al., 2004 U65992
GluR4d (chicken only) U65993
GluR4s (chicken only) U65991
IL1RL1/ST2 interleukin 1 receptor-like 1 ST2 D12763 AB041737
ST2L AB012701 AB041738
ST2V AB029084 Not found
ST2LV (chicken only) AB041739
MAPK9 c-Jun N-terminal kinase 2 JNK2-alpha1 NM_139068 NM_205095
JNK2-alpha2 NM_002752 Not found
JNK2-beta1 NM_139069 Not found
JNK2-beta2 NM_139070 Not found
MAPT microtubule-associated protein tau MAPT variant 1 NM_016835 Yoshida and Goedert, 2002
MAPT variant 2 NM_005910 Yoshida and Goedert, 2002
MAPT variant 3 NM_016834 Yoshida and Goedert, 2002
MAPT variant 4 NM_016841 Yoshida and Goedert, 2002
MYB c-myb c-myb+9a Dasgupta and Reddy, 1989 Schuur et al., 1993
c-myb (−9a) NM_005375 NM_205306
NFASC neurofascin NFASC variant 1 NM_001005388 XM_417958
NFASC variant 2 NM_015090 Hassel et al., 1997
NFASC variant 3 NM_001005389 Hassel et al, 1997
NGF nerve growth factor Multiple multiple
PDE5A cGMP-binding, cGMP-specific 3′, 5′-cyclic nucleotide phosphodiesterase PDE5A variant 1 NM_001083 Giordano et al., 2004
PDE5A variant 2 NM_033430 Giordano et al., 2004
PDE5A variant 3 NM_033437 Giordano et al., 2004
PSAP prosaposin hPSAP0 NM_002778 Cohen et al., 2004
hPSAP9 NM_001042465 Cohen et al., 2004
hPSAP6 NM_001042466 Absent in chicken, Cohen et al., 2004
SRC Src Src NM_005417 XM_425696
neuronal Src-NI Pyper and Bolen, 1990 Martinez et al., 1987
neuronal Src-NII Pyper and Bolen, 1990 Not found
SST somatostatin SS1/PSS1 NM_001048 NM_205336
SS2/PSS2 Trabucchi et al., 2003 AY221945
TERT telomerase reverse transcriptase multiple multiple
TPM2 beta-tropomyosin TPM2–6a NM_213674 Libri et al., 1991
TPM2–6b NM_003289 Libri et al., 1991

NGF

Nerve growth factor is a member of the neurotrophin family which consists of functionally and structurally related small basic proteins of ~ 120 amino acids. Analysis of the chicken NGF gene and transcripts reveals five alternatively spliced 5′ non-coding exons grouped into two clusters, followed by the coding exon 4. The first cluster consists of three leader exons (1a, 1b, 1c) and the second cluster contains exons 2 and 3. Differential usage of leader and internal exons, alternative transcription start sites, alternative donor and acceptor sites generates at least 21 different transcripts (Bertaux et al., 2004). Comparison of the 6-exon chicken NGF with the 3-exon human NGFB gene reveals significant nucleotide alignment of chicken NGF exons 2 and 4 with human NGFB exons 2 and 3, with the chicken and human proteins showing 87% identity. There is no equivalent to chicken exon 3 in either the human or mouse genome (Bertaux et al., 2004).

TERT

Telomerase reverse transcriptase (TERT) adds repeats to telomeres during DNA replication to maintain telomere length. Recent evidence suggests that TERT may have additional roles in the regulation of cell growth and tumor formation. In humans, four insertion variants of TERT have been identified, as well as three deletion variants and various combinations thereof (Kilian et al., 1997; Ulaner et al., 1998; Hisatomi et al., 2003). Nineteen variants have been identified in chicken, involving exon skipping, intron retention, and alternative usage of splice donor and acceptor sites (Chang and Delany, 2006). The number of variants ranged from ten in adult chicken liver to 13 in chicken embryo fibroblasts. One of the variants, found in all tissues examined, is predicted to generate a truncated product lacking telomerase activity.

Somatostatin

The neuropeptide somatostatin was first isolated based on its ability to inhibit growth hormone secretion from the anterior pituitary (Brazeau et al., 1973). Two biologically active forms of somatostatin (SS1, SS2) generated through alternative splicing of precursor somatostatin (PSS1, PSS2) have been identified in mammals, amphibians, fish and chicken (reviewed in Trabucchi et al., 2003). Whereas PSS1 is widely expressed in the central nervous system, where it functions as a neurotransmitter and a neuromodulator, PSS2 is only expressed in specific regions of the brain.

JNK2-α1

c-Jun N-terminal kinases (JNK) constitute a subgroup of the mitogen-activated protein kinases (MAPK). Alternatively spliced isoforms of mammalian JNK2 which differ in their carboxy-terminus have been identified; however, extensive analysis of JNK2 in chicken revealed only the C-terminus truncated subtype α1 (Ishikawa et al., 1997). JNK2-α1 is predominantly expressed in the developing chick brain, suggesting a role in neuronal cell formation.

GRIA4/AMPA4

AMPA receptors, which mediate rapid excitatory synaptic transmission in the vertebrate CNS, are pentameric assemblies of four glutamate receptors (GIuR1–4). The type and amount of GluR incorporated into an AMPA receptor determines the receptor’s properties. Each GluR, including GluR4 (GRIA4), exists in two alternatively spliced forms, called flip and flop that are evolutionarily conserved in chicken and mammals (Ravindranathan et al., 1997). In addition, a C-terminal splice variant called GluR4c, which also exists in flip and flop forms, has been described in mammals and chicken (Ravindranathan et al., 1997; Kawahara et al., 2004). Two additional alternatively spliced variants, GluR4d with a 184-bp fragment inserted at the 4c splice site, and GluR4s, a shortened version of GluR4 that lacks the 4th transmembrane and flip/flop domains have been described in chicken (Ravindranathan et al., 1997). Mammalian GluR4d and Glu4s have not yet been reported.

FGF8, FGFR2

Fibroblast growth factor 8 (FGF8) is a signaling molecule involved in cell proliferation and differentiation. The human genomic structure predicts four alternatively spliced isoforms, three of which have been identified (Ghosh et al., 1996) while the mouse genomic structure predicts eight alternatively spliced isoforms, seven of which have been identified (reviewed in Haworth et al., 2005). Extensive analysis of chicken FGF8 has revealed only two isoforms, equivalent to mouse FGF8-a and FGF8-b (Haworth et al., 2005).

Alternative splicing of FGF receptors (FGFR) results in different ligand binding properties and alterations in biological function. The FGFR2 gene encodes two well-characterized alternatively spliced products: FGFR2b and FGFR2c. These two isoforms are identical except for a 49 amino acid sequence in the extracellular region (Ornitz and Marie, 2002). These isoforms exhibit different ligand-binding properties as well as distinct expression patterns in the developing embryo (Ornitz and Marie, 2002). Both FGFR2 isoforms have been described in chicken (Havens et al., 2006).

ST2

The ST2 gene encodes receptor-like molecules that are similar to interleukin-1 receptors. The chicken, mouse, rat and human ST2 genes consist of 13 exons and two promoters followed by two non-coding exons 1a and 1b. Three ST2 variants, generated by alternative splicing, have been described in humans (Tominaga et al., 1999; Li et al., 2000). Analysis of chicken ST2 revealed three variants, ST2 and ST2L previously described in humans, as well as a novel variant called ST2LV (Iwahana et al., 2004).

PDE5

Three splice variants of cGMP-binding cGMP-specific phosphodiesterase (PDE5) have been described in humans: PDE5A1, A2, A3 (Lin et al., 2000). All three variants have recently been described in chicken dorsal root ganglia (Giordano et al., 2004).

Drebrin

Drebrins (developmentally regulated brain proteins, DBN1) are actin binding proteins believed to play a role in the formation of actin filaments in dendritic spines. Three isoforms, generated by alternative splicing, have been identified in chicken brain: two embryonic types (E1 and E2) and an adult type (A) (Kojima et al, 1993). Two DBN1 isoforms have been reported in human (1a (E2) and 1b (A)), three in mice (A, A2, E2) and two in rat (A and E) (Chew et al., 2005).

Neurofascin

Neurofascin (NFASC) is a neural cell adhesion molecule implicated in cell adhesion, cell migration and neurite outgrowth. The chicken NFASC gene has 33 exons, eight of which have been shown to undergo alternative splicing. Analysis of 138 independent NFASC cDNAs derived from chick embryonic brain at day 6 and day 16 of incubation led to the conclusion that there are at least 50 different NFASC splice products expressed at different developmental stages in chicken (Hassel et al., 1997). A large number of different NFASC splice products have also been found in humans (http://www.ensembl.org/index.html).

Prosaponin

Prosaponin is the precursor of the lysosomal activator molecules saponins A, B, C and D. The prosaponin gene has 15 exons, one of which (the 9-bp exon 8) undergoes alternative splicing. Three prosaponin mRNAs have been identified in human and mouse, containing 0,6 or 9 bp of exon 8. In chicken, there are only two prosaponin variants, containing 0 or 9 bp of exon 8 (Cohen et al., 2004). This suggests that the 6-bp exon 8 splice variant may not be biologically important, in keeping with the low levels of this splice product (Cohen et al, 2004).

Tau

Tau is a microtubule-associated protein expressed in brain. Alternative splicing of a single tau (MAPT) gene generates five isoforms in adult chicken brain (Yoshida and Goedert, 2002) and six isoforms in human brain (Goedert et al., 1989; Andreadis et al., 1992). The chicken and human tau isoforms differ by the number of repeated microtubule-binding domains that they contain. Isoforms with the highest number of repeats are thus better at promoting microtubule assembly. Whereas human tau adult isoforms consist of three or four repeats, two of the five adult chicken tau isoforms consist of five repeats. Both human and chicken express an additional isoform, a three-repeat tau without an insert in the N-terminus, which predominates at early developmental stages (Yoshida and Goedert, 2002).

Casein kinase 1

Casein kinase 1 is a family of serine/threonine protein kinases shown to phosphorylate acidic peptides in a variety of substrates. At least two members of the family, α and γ3, undergo alternative splicing. There are four splice variants of chicken CKIα which differ by the presence or absence of two peptides, a 28 amino acid L insert containing a nuclear localization signal and a 12 amino acid S insert with no known motifs (Green and Bennett, 1998; Fu et al., 2001). The four splice variants, CKIα, CKIαS, CKIαL and CKIαLS have specific patterns of expression in chicken tissues (Green and Bennett, 1998). Three of the four splice variants (CKIα, CKIαL, CKIαS) have been reported in mammals (Zhang et al., 1996; Yong et al., 2000).

Myb

Two alternatively spliced forms of c-myb have been reported in chicken and mammals. These splice variants differ by the inclusion/exclusion of an exon (9A) encoding 120 (chicken) or 121 (human, mouse) amino acids (Dasgupta and Reddy, 1989; Schuur et al., 1993). This exon is not present in v-myb. Functional analysis of Myb(+9a) suggests that it has stronger transactivation ability compared to Myb (−9a) although both forms are equally efficient at transforming primary chicken hematopoietic cells (Woo et al., 1998). Alternative splicing of exon 9A has also been described for B-myb (Kamano et al., 1995). Interestingly, B-myb(−9a) lacks transactivation activity as part of the transactivation domain of B-myb is found in exon 9a (Horstmann et al, 2000).

c-Src

Cellular Src is a membrane-associated tyrosine protein kinase with oncogenic potential. c-Src levels are elevated in neuronal tissue suggesting a role in neuronal differentiation. A neuron-specific isoform of c-Src (NI) which contains an 18-bp exon inserted between exons 3 and 4 has been identified in mammals and chicken (Lynch et al., 1986; Levy et al, 1987). Exons 3 and 4 encode part of the domain involved in substrate interaction and regulation of c-Src activity. A second neuronal-specific isoform of c-Src (NII) has been identified in human brain, consisting of a 33-bp exon inserted between exons 3 and 4 (Pyper and Bolen, 1990). The NII isoform has not been reported in chicken. Alternative splicing of c-Src, involving different splice products, has also been described in chicken skeletal muscle cells (Dorai and Wang, 1990).

Alternative splicing in the top fifty genes listed under GeneCards

As a third strategy, we compared alternative splicing in the 50 most frequently accessed genes based on the GeneCards database (http://www.genecards.org/). Table 3 indicates the number of alternatively spliced products listed in the GeneCards and Ensembl databases for the human genes, compared to the number of alternatively spliced products listed in the Ensembl database for the chicken orthologues.

Table 3.

Alternative splicing in 50 commonly accessed genes

Gene symbol Gene name Number of splicing variants
Conserved transcriptsb
Human
Chicken
GeneCardsa Ensembl Ensemblb
BRCA1 breast cancer 1, early onset 9 8 1 0/1
TP53 tumor protein p53 (Li-Fraumeni syndrome) 4 2 NA NA
ALDH2 aldehyde dehydrogenase 2 family (mitochondrial) 4 1 1 0/1
COL2A1 collagen, type II, alpha 1 6 4 1 0/1
CFTR cystic fibrosis transmembrane conductance regulator, ATP-binding cassette (sub-family C, member 7) 1 1 4 0/4
BTK Bruton agammaglobulinemia tyrosine kinase 6 7 2 0/2
A2M alpha-2-macroglobulin 1 1 NA NA
BRCA2 breast cancer 2, early onset 1 2 1 0/1
ASS argininosuccinate synthetase 11 10 NA NA
EGFR epidermal growth factor receptor 1 3 3 0/3
IGF1 insulin-like growth factor 1 1 2 1 1/1
BCL2 B-cell CLL/lymphoma 2 2 1 1 1/1
APP amyloid beta (A4) precursor protein 1 7 1 1/1
APOE apolipoprotein E 4 1 NA NA
ATM ataxia telangiectasia mutated 1 2 1 0/1
HEXA hexosaminidase A 11 3 1 1/1
AKT1 v-akt murine thymoma viral oncogene homolog 1 8 1 1 0/1
HBB hemoglobin, beta 1 5 3 NA
RBI retinoblastoma 1 1 3 2 0/2
JAK2 Janus kinase 2 1 1 1 0/1
AR androgen receptor 1 2 1 0/1
BAX BCL2-associated X protein 3 1 NA NA
JAK3 Janus kinase 3 1 4 NA NA
PTEN phosphatase and tensin homolog 1 2 1 1/1
GBA glucosidase, beta; acid 10 14 (12 without translation) NA NA
HFE hemochromatosis 11 9 0
SOD1 superoxide dismutase 1, soluble 7 1 1 1/1
CREBBP CREB binding protein 1 3 1 0/1
INSR insulin receptor 1 2 1 0/1
ACE angiotensin I converting enzyme 12 2 2 0/2
ADRB2 adrenergic, beta-2-, receptor, surface 1 1 1 1/1
TNF tumor necrosis factor (TNF superfamily, member 2) 1 4 NA NA
JUN v-jun sarcoma virus 17 oncogene homolog (avian) 1 1 1 0/1
PSAP prosaposin 8 2 2 1/2
ATBF1 AT-binding transcription factor 1 1 1 1 0/1
NFKB1 nuclear factor of kappa light polypeptide gene enhancer in B-cells 1 (P105) 4 1 1 0/1
ACHE acetylcholinesterase (YT blood group) 10 3 NA NA
VEGF vascular endothelial growth factor 6 9 (2 without translation) 5 2/5
APC adenomatosis polyposis coli 1 1 1 0/1
ABL1 v-abl Abelson murine leukemia viral oncogene homolog 1 1 2 1 1/1
AHCY S-adenosylhomocysteine hydrolase 9 2 1 1/1
ABCB1 ATP-binding cassette, sub-family B (MDR/TAP), member 1 3 1 2 0/2
ADRB3 adrenergic, beta-3-, receptor 1 1 1 0/1
CASP8 caspase 8, apoptosis-related cysteine peptidase 14 9 1 0/1
FGFR3 fibroblast growth factor receptor 3 1 5 1 0/1
MTHFR 5,10-methylenetetrahydrofolate reductase (NADPH) 8 5 1 0/1
HEXB hexosaminidase B (beta polypeptide) 1 2 1 0/1
DMD dystrophin (muscular dystrophy, Duchenne and Becker types) 1 19 1 0/1
PPARG peroxisome proliferative activated receptor, gamma 7 2 1 0/1
ARSA arylsulfalase A 8 2 1 1/1
a

‘Hot genes’ are defined genes that have been most frequently requested at the GeneCards home site (http://www.genecards.org/) since Jan 1, 2006 sorted by frequency of access.

b

‘NA’ denotes information not available.

As shown in Table 3, alternative splicing is much more commonly reported for human genes than for their chicken counterparts, likely reflecting the extent of work that has been carried out on analysis of human versus chicken genes to date. Of the 50 genes analysed, only one gene, CFTR, had more alternatively spliced variants in chicken compared to human. Alternatively spliced products were reported in 9/50 chicken genes compared to 42/50 human genes, although there was significant variation in the number of alternatively spliced products listed in the GeneCards and Ensembl databases for many human genes. Conservation of splice variants between chicken and human was observed for 12 genes which had more than one human splice variant; however 11 of these genes had only one reported chicken transcript, suggesting that the chicken transcript represented a main or constitutive splice product. Two conserved splice variants in human and chicken were identified for the VEGF gene.

Conclusions

A survey of 50 commonly accessed genes reveals little conservation in alternative splicing between chicken and human. However, rather than denoting a penury of splicing variants in chicken, this likely reflects the fact that mRNA sequence analysis is still at an early stage in chicken. Specific analysis of 16 genes known to undergo alternative splicing in chicken reveals as many alternatively spliced products in chicken compared to mammalian species, with a total of 37 splice products in chicken and 38 in humans. Furthermore, for these well-characterized chicken genes, the same alternatively spliced products are commonly found in both chicken and mammalian species (32 out of 38 human splice products were conserved in chicken), suggesting evolutionary conservation of splice variant function. Perhaps the most informative genes are those selected on the basis of expression in the well-conserved retina with well-documented alternative splicing in mammals. Of the 65 mammalian splice variants shown in Table 1, 42 could be documented as conserved in chicken. Although this study is limited in scope because of a general lack of mRNA data in chicken, our results support a high frequency of alternative splicing events in chicken, comparable to that observed in mammals. Brett et al. (2002) have compared the number of alternative splicing events in seven eukaryotic species for which sufficient mRNA/EST data were available (human, mouse, rat, cow, fly, worm, plant). When differences in mRNA and EST coverage were taken into consideration, these investigators found that the amount of alternative splicing was similar in all seven organisms, with 10% of mRNAs having alternatively spliced forms. These results suggest that the complexity between higher and lower eukaryotes may not be reflected in the number of alternative splice events, but rather in the nature of alternative splice events.

Acknowledgments

Supported by the Canadian Institutes for Health Research.

References

  1. Andreadis A, Brown WM, Kosik KS. Structure and novel exons of the human tau gene. Biochemistry. 1992;31:10626–10633. doi: 10.1021/bi00158a027. [DOI] [PubMed] [Google Scholar]
  2. Azuma N, Tadokoro K, Asaka A, Yamada M, Yamaguchi Y, et al. The Pax6 isoform bearing an alternative spliced exon promotes the development of the neural retinal structure. Hum Mol Genet. 2005;14:735–745. doi: 10.1093/hmg/ddi069. [DOI] [PubMed] [Google Scholar]
  3. Bar I, Tissir F, Lambert de RC, De BO, Goffinet AM. The gene encoding disabled-1 (DAM), the intracellular adaptor of the Reelin pathway, reveals unusual complexity in human and mouse. J Biol Chem. 2003;278:5802–5812. doi: 10.1074/jbc.M207178200. [DOI] [PubMed] [Google Scholar]
  4. Bates DO, Cui TG, Doughty JM, Winkler M, Sugiono M, et al. VEGF165b, an inhibitory splice variant of vascular endothelial growth factor, is down-regulated in renal cell carcinoma. Cancer Res. 2002;62:4123–4131. [PubMed] [Google Scholar]
  5. Benke D, Michel C, Mohler H. Structure of GABAB receptors in rat retina. J Recept Signal Transduct Res. 2002;22:253–266. doi: 10.1081/rrs-120014600. [DOI] [PubMed] [Google Scholar]
  6. Bertaux O, Toselli-Mollereau E, Auffray C, Devignes MD. Alternative usage of 5′ exons in the chicken nerve growth factor gene: refined characterization of a weakly expressed gene. Gene. 2004;334:83–97. doi: 10.1016/j.gene.2004.02.052. [DOI] [PubMed] [Google Scholar]
  7. Billinton A, Ige AO, Bolam JP, White JH, Marshall FH, Emson PC. Advances in the molecular understanding of GABA(B) receptors. Trends Neurosci. 2001;24:277–282. doi: 10.1016/s0166-2236(00)01815-4. [DOI] [PubMed] [Google Scholar]
  8. Black DL. Mechanisms of alternative pre-messenger RNA splicing. Annu Rev Biochem. 2003;72:291–336. doi: 10.1146/annurev.biochem.72.121801.161720. [DOI] [PubMed] [Google Scholar]
  9. Boue S, Letunic I, Bork P. Alternative splicing and evolution. Bioessays. 2003;25:1031–1034. doi: 10.1002/bies.10371. [DOI] [PubMed] [Google Scholar]
  10. Brazeau P, Vale W, Burgus R, Ling N, Butcher M, et al. Hypothalamic polypeptide that inhibits the secretion of immunoreactive pituitary growth hormone. Science. 1973;179:77–79. doi: 10.1126/science.179.4068.77. [DOI] [PubMed] [Google Scholar]
  11. Brett D, Pospisil H, Valcarcel J, Reich J, Bork P. Alternative splicing and genome complexity. Nat Genet. 2002;30:29–30. doi: 10.1038/ng803. [DOI] [PubMed] [Google Scholar]
  12. Burns ME, Mendez A, Chen CK, Almuete A, Quillinan N, et al. Deactivation of phosphorylated and nonphosphorylated rhodopsin by arrestin splice variants. J Neurosci. 2006;26:1036–1044. doi: 10.1523/JNEUROSCI.3301-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Chang H, Delany ME. Complicated RNA splicing of chicken telomerase reverse transcriptase revealed by profiling cells both positive and negative for telomerase activity. Gene. 2006;379:33–39. doi: 10.1016/j.gene.2006.04.021. [DOI] [PubMed] [Google Scholar]
  14. Chen F, Figueroa DJ, Marmorstein AD, Zhang Q, Petrukhin K, et al. Retina-specific nuclear receptor: A potential regulator of cellular retinal-dehyde-binding protein expressed in retinal pigment epithelium and Muller glial cells. Proc Natl Acad Sci USA. 1999;96:15149–15154. doi: 10.1073/pnas.96.26.15149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Chew CS, Okamoto CT, Chen X, Thomas R. Drebrin E2 is differentially expressed and phosphorylated in parietal cells in the gastric mucosa. Am J Physiol Gastrointest Liver Physiol. 2005;289:G320–G331. doi: 10.1152/ajpgi.00002.2005. [DOI] [PubMed] [Google Scholar]
  16. Cohen T, Ravid L, Altman N, Madar-Shapiro L, Fein A, et al. Conservation of expression and alternative splicing in the prosaposin gene. Brain Res Mol Brain Res. 2004;129:8–19. doi: 10.1016/j.molbrainres.2004.06.027. [DOI] [PubMed] [Google Scholar]
  17. Conn PJ, Pin JP. Pharmacology and functions of metabotropic glutamate receptors. Annu Rev Pharmacol Toxicol. 1997;37:205–237. doi: 10.1146/annurev.pharmtox.37.1.205. [DOI] [PubMed] [Google Scholar]
  18. Croft L, Schandorff S, Clark F, Burrage K, Arctander P, Mattick JS. ISIS, the intron information system, reveals the high frequency of alternative splicing in the human genome. Nat Genet. 2000;24:340–341. doi: 10.1038/74153. [DOI] [PubMed] [Google Scholar]
  19. Dasgupta P, Reddy EP. Identification of alternatively spliced transcripts for human c-myb: molecular cloning and sequence analysis of human c-myb exon 9A sequences. Oncogene. 1989;4:1419–1423. [PubMed] [Google Scholar]
  20. Dho SE, French MB, Woods SA, McGlade CJ. Characterization of four mammalian numb protein isoforms. Identification of cytoplasmic and membrane-associated variants of the phosphotyrosine binding domain. J Biol Chem. 1999;274:33097–33104. doi: 10.1074/jbc.274.46.33097. [DOI] [PubMed] [Google Scholar]
  21. Dooley CM, James J, Jane MC, Ahmad I. Involvement of numb in vertebrate retinal development: evidence for multiple roles of numb in neural differentiation and maturation. J Neurobiol. 2003;54:313–325. doi: 10.1002/neu.10176. [DOI] [PubMed] [Google Scholar]
  22. Dorai T, Wang LH. An alternative non-tyrosine protein kinase product of the c-src gene in chicken skeletal muscle. Mol Cell Biol. 1990;10:4068–4079. doi: 10.1128/mcb.10.8.4068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Ferrara N, Davis-Smyth T. The biology of vascular endothelial growth factor. Endocr Rev. 1997;18:4–25. doi: 10.1210/edrv.18.1.0287. [DOI] [PubMed] [Google Scholar]
  24. Francesconi A, Duvoisin RM. Alternative splicing unmasks dendritic and axonal targeting signals in metabotropic glutamate receptor 1. J Neurosci. 2002;22:2196–2205. doi: 10.1523/JNEUROSCI.22-06-02196.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Fu Z, Chakraborti T, Morse S, Bennett GS, Shaw G. Four casein kinase I isoforms are differentially partitioned between nucleus and cytoplasm. Exp Cell Res. 2001;269:275–286. doi: 10.1006/excr.2001.5324. [DOI] [PubMed] [Google Scholar]
  26. Ghosh AK, Shankar DB, Shackleford GM, Wu K, T’Ang A, et al. Molecular cloning and characterization of human FGF8 alternative messenger RNA forms. Cell Growth Differ. 1996;7:1425–1434. [PubMed] [Google Scholar]
  27. Giordano D, Giorgi M, Tata AM, Modica A, Augusti-Tocco G. Expression of PDE5 splice variants during ontogenesis of chick dorsal root ganglia. J Neurosci Res. 2004;78:815–823. doi: 10.1002/jnr.20337. [DOI] [PubMed] [Google Scholar]
  28. Glencorse TA, Bateson AN, Darlison MG. Differential localization of two alternatively spliced GABAA receptor gamma2-subunit mRNAs in the Chick Brain. Eur J Neurosci. 1992;4:271–277. doi: 10.1111/j.1460-9568.1992.tb00874.x. [DOI] [PubMed] [Google Scholar]
  29. Godbout R. High levels of aldehyde dehydrogenase transcripts in the undifferentiated chick retina. Exp Eye Res. 1992;54:297–305. doi: 10.1016/s0014-4835(05)80219-2. [DOI] [PubMed] [Google Scholar]
  30. Godbout R, Monckton EA. Differential regulation of the aldehyde dehydrogenase 1 gene in embryonic chick retina and liver. J Biol Chem. 2001;276:32896–32904. doi: 10.1074/jbc.M104372200. [DOI] [PubMed] [Google Scholar]
  31. Godbout R, Packer M, Poppema S, Dabbagh L. Localization of cytosolic aldehyde dehydrogenase in the developing chick retina: in situ hybridization and immunohistochemical analyses. Dev Dyn. 1996;205:319–331. doi: 10.1002/(SICI)1097-0177(199603)205:3<319::AID-AJA11>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
  32. Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA. Multiple isoforms of human mirotubule-associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer’s disease. Neuron. 1989;3:519–526. doi: 10.1016/0896-6273(89)90210-9. [DOI] [PubMed] [Google Scholar]
  33. Green CL, Bennett GS. Identification of four alternatively spliced isoforms of chicken casein kinase I alpha that are all expressed in diverse cell types. Gene. 1998;216:189–195. doi: 10.1016/s0378-1119(98)00291-1. [DOI] [PubMed] [Google Scholar]
  34. Haber DA, Sohn RL, Buckler AJ, Pelletier J, Call KM, Housman DE. Alternative splicing and genomic structure of the Wilms tumor gene WT1. Proc Natl Acad Sci USA. 1991;88:9618–9622. doi: 10.1073/pnas.88.21.9618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Harvey RJ, Chinchetru MA, Darlison MG. Alternative splicing of a 51-nucleolide exon that encodes a putative protein kinase C phosphorylation site generates two forms of the chicken gamma-aminobutyric acidA receptor beta 2 subunit. J Neurochem. 1994;62:10–16. doi: 10.1046/j.1471-4159.1994.62010010.x. [DOI] [PubMed] [Google Scholar]
  36. Hassel B, Rathjen FG, Volkmer H. Organization of the neurofascin gene and analysis of developmentally regulated alternative splicing. J Biol Chem. 1997;272:28742–28749. doi: 10.1074/jbc.272.45.28742. [DOI] [PubMed] [Google Scholar]
  37. Havens BA, Rodgers B, Mina M. Tissue-specific expression of Fgfr2b and Fgfr2c isoforms, Fgf10 and Fgf9 in the developing chick mandible. Arch Oral Biol. 2006;51:134–145. doi: 10.1016/j.archoralbio.2005.06.011. [DOI] [PubMed] [Google Scholar]
  38. Haworth KE, Healy C, Sharpe PT. Characterisation of the genomic structure of chick Fgf8. DNA Seq. 2005;16:180–186. doi: 10.1080/10425170500069973. [DOI] [PubMed] [Google Scholar]
  39. Hayashi T, Huang J, Deeb SS. RINX(VSX1), a novel homeobox gene expressed in the inner nuclear layer of the adult retina. Genomics. 2000;67:128–139. doi: 10.1006/geno.2000.6248. [DOI] [PubMed] [Google Scholar]
  40. Hisatomi H, Ohyashiki K, Ohyashiki JH, Nagao K, Kanamaru T, et al. Expression profile of a gamma-deletion variant of the human telomerase reverse transcriptase gene. Neoplasia. 2003;5:193–197. doi: 10.1016/S1476-5586(03)80051-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Horstmann S, Ferrari S, Klempnauer KH. An alternatively spliced isoform of B-Myb is a transcriptional inhibitor. Oncogene. 2000;19:5428–5434. doi: 10.1038/sj.onc.1203937. [DOI] [PubMed] [Google Scholar]
  42. Huang W, Lu N, Eberspaecher H, de Crombrugghe B. A new long form of c-Maf cooperates with Sox9 to activate the type II collagen gene. J Biol Chem. 2002;277:50668–50675. doi: 10.1074/jbc.M206544200. [DOI] [PubMed] [Google Scholar]
  43. Ishikawa T, Nakada-Moriya Y, Ando C, Tanda N, Nishida S, et al. Expression of the JNK2-alpha 1 gene in the developing chick brain. Biochem Biophys Res Commun. 1997;234:489–492. doi: 10.1006/bbrc.1997.6140. [DOI] [PubMed] [Google Scholar]
  44. Isomoto S, Kaibara M, Sakurai-Yamashita Y, Nagayama Y, Uezono Y, et al. Cloning and tissue distribution of novel splice variants of the rat GABAB receptor. Biochem Biophys Res Commun. 1998;253:10–15. doi: 10.1006/bbrc.1998.9706. [DOI] [PubMed] [Google Scholar]
  45. Iwahana H, Hayakawa M, Kuroiwa K, Tago K, Yanagisawa K, et al. Molecular cloning of the chicken ST2 gene and a novel variant form of the ST2 gene product, ST2LV. Biochim Biophys Acta. 2004;1681:1–14. doi: 10.1016/j.bbaexp.2004.08.013. [DOI] [PubMed] [Google Scholar]
  46. Johnson JM, Castle J, Garrett-Engele P, Kan Z, Loerch PM, et al. Genome-wide survey of human alternative pre-mRNA splicing with exon junction microarrays. Science. 2003;302:2141–2144. doi: 10.1126/science.1090100. [DOI] [PubMed] [Google Scholar]
  47. Kamano H, Burk B, Noben-Trauth K, Klempnauer KH. Differential splicing of the mouse B-myb gene. Oncogene. 1995;11:2575–2582. [PubMed] [Google Scholar]
  48. Kan Z, Garrett-Engele PW, Johnson JM, Castle JC. Evolutionarily conserved and diverged alternative splicing events show different expression and functional profiles. Nucleic Acids Res. 2005;33:5659–5666. doi: 10.1093/nar/gki834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Katyal S, Godbout R. Alternative splicing modulates Disabled-1 (Dab1) function in the developing chick retina. EMBO J. 2004;23:1878–1888. doi: 10.1038/sj.emboj.7600185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Kawahara Y, Ito K, Sun H, Ito M, Kanazawa I, Kwak S. GluR4c, an alternative splicing isoform of GluR4, is abundantly expressed in the adult human brain. Brain Res Mol Brain Res. 2004;127:150–155. doi: 10.1016/j.molbrainres.2004.05.020. [DOI] [PubMed] [Google Scholar]
  51. Kawakami K, Ohto H, Takizawa T, Saito T. Identification and expression of six family genes in mouse retina. FEBS Lett. 1996;393:259–263. doi: 10.1016/0014-5793(96)00899-x. [DOI] [PubMed] [Google Scholar]
  52. Kilian A, Bowtell DD, Abud HE, Hime GR, Venter DJ, et al. Isolation of a candidate human telomerase catalytic subunit gene, which reveals complex splicing patterns in different cell types. Hum Mol Genet. 1997;6:2011–2019. doi: 10.1093/hmg/6.12.2011. [DOI] [PubMed] [Google Scholar]
  53. Kirsch J, Kroger S. Postsynaptic anchoring of receptors: A cellular approach to neuronal and muscular sensitivity. J Neuroscientist. 1996;2:100–108. [Google Scholar]
  54. Kojima N, Shirao T, Obata K. Molecular cloning of a developmentally regulated brain protein, chicken drebrin A and its expression by alternative splicing of the drebrin gene. Brain Res Mol Brain Res. 1993;19:101–114. doi: 10.1016/0169-328x(93)90154-h. [DOI] [PubMed] [Google Scholar]
  55. Kondrashov FA, Koonin EV. Evolution of alternative splicing: deletions, insertions and origin of functional parts of proteins from intron sequences. Trends Genet. 2003;19:115–119. doi: 10.1016/S0168-9525(02)00029-X. [DOI] [PubMed] [Google Scholar]
  56. Kreutz MR, Bockers TM, Bockmann J, Seidenbecher CI, Kracht B, et al. Axonal injury alters alternative splicing of the retinal NR1 receptor: the preferential expression of the NR1b isoform is crucial for retinal ganglion cell survival. J Neurosci. 1998;18:8278–8291. doi: 10.1523/JNEUROSCI.18-20-08278.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Kroger S. Differential distribution of agrin isoforms in the developing and adult avian retina. Mol Cell Neurosci. 1997;10:149–161. doi: 10.1006/mcne.1997.0651. [DOI] [PubMed] [Google Scholar]
  58. Lareau LF, Green RE, Bhatnagar RS, Brenner SE. The evolving roles of alternative splicing. Curr Opin Struct Biol. 2004;14:273–282. doi: 10.1016/j.sbi.2004.05.002. [DOI] [PubMed] [Google Scholar]
  59. Laurie DJ, Boddeke HW, Hiltscher R, Sommer B. HmGluld, a novel splice variant of the human type I metabotropic glutamate receptor. Eur J Pharmacol. 1996;296:R1–R3. doi: 10.1016/0014-2999(95)00868-3. [DOI] [PubMed] [Google Scholar]
  60. Levy JB, Dorai T, Wang LH, Brugge JS. The structurally distinct form of pp60c-src detected in neuronal cells is encoded by a unique c-src mRNA. Mol Cell Biol. 1987;7:4142–4145. doi: 10.1128/mcb.7.11.4142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Li H, Tago K, Io K, Kuroiwa K, Arai T, et al. The cloning and nucleotide sequence of human ST2L cDNA. Genomics. 2000;67:284–290. doi: 10.1006/geno.2000.6269. [DOI] [PubMed] [Google Scholar]
  62. Libri D, Piseri A, Fiszman MY. Tissue-specific splicing in vivo of the beta-tropomyosin gene: dependence on an RNA secondary structure. Science. 1991;252:1842–1845. doi: 10.1126/science.2063196. [DOI] [PubMed] [Google Scholar]
  63. Lin CS, Lau A, Tu R, Lue TF. Expression of three isoforms of cGMP-binding cGMP-specific phosphodiesterase (PDE5) in human penile cavernosum. Biochem Biophys Res Commun. 2000;268:628–635. doi: 10.1006/bbrc.2000.2187. [DOI] [PubMed] [Google Scholar]
  64. Lynch SA, Brugge JS, Levine JM. Induction of altered c-src product during neural differentiation of embryonal carcinoma cells. Science. 1986;234:873–876. doi: 10.1126/science.3095923. [DOI] [PubMed] [Google Scholar]
  65. Makoff AJ, Phillips T, Pilling C, Emson P. Expression of a novel splice variant of human mGluR1 in the cerebellum. Neuroreport. 1997;8:2943–2947. doi: 10.1097/00001756-199709080-00027. [DOI] [PubMed] [Google Scholar]
  66. Martinez R, Mathey-Prevot B, Bernards A, Baltimore D. Neuronal pp60c-src contains a six-amino acid insertion relative to its non-neuronal counterpart. Science. 1987;237:411–415. doi: 10.1126/science.2440106. [DOI] [PubMed] [Google Scholar]
  67. McCaffery P, Tempst P, Lara G, Drager UC. Aldehyde dehydrogenase is a positional marker in the retina. Development. 1991;112:693–702. doi: 10.1242/dev.112.3.693. [DOI] [PubMed] [Google Scholar]
  68. McCaffery P, Lee MO, Wagner MA, Sladek NE, Drager UC. Asymmetrical retinoic acid synthesis in the dorsoventral axis of the retina. Development. 1992;115:371–382. doi: 10.1242/dev.115.2.371. [DOI] [PubMed] [Google Scholar]
  69. McKinley DD, Lennon DJ, Carter DB. Cloning, sequence analysis and expression of two forms of mRNA coding for the human beta 2 subunit of the GABAA receptor. Brain Res Mol Brain Res. 1995;28:175–179. doi: 10.1016/0169-328x(94)00228-7. [DOI] [PubMed] [Google Scholar]
  70. Meindl A, Dry K, Herrmann K, Manson F, Ciccodicola A, et al. A gene (RPGR) with homology to the RCC1 guanine nucleotide exchange factor is mutated in X-linked retinitis pigmentosa (RP3) Nat Genet. 1996;13:35–42. doi: 10.1038/ng0596-35. [DOI] [PubMed] [Google Scholar]
  71. Modrek B, Lee C. A genomic view of alternative splicing. Nat Genet. 2002;30:13–19. doi: 10.1038/ng0102-13. [DOI] [PubMed] [Google Scholar]
  72. Modrek B, Lee CJ. Alternative splicing in the human, mouse and rat genomes is associated with an increased frequency of exon creation and/or loss. Nat Genet. 2003;34:177–180. doi: 10.1038/ng1159. [DOI] [PubMed] [Google Scholar]
  73. Nishida A, Furukawa A, Koike C, Tano Y, Aizawa S, Matsuo I, Furukawa T. Otx2 homeobox gene controls retinal photoreceptor cell fate and pineal gland development. Nat Neurosci. 2003;6:1255–1263. doi: 10.1038/nn1155. [DOI] [PubMed] [Google Scholar]
  74. O’Connor LT, Lauterborn JC, Gall CM, Smith MA. Localization and alternative splicing of agrin mRNA in adult rat brain: transcripts encoding isoforms that aggregate acetylcholine receptors are not restricted to cholinergic regions. J Neurosci. 1994;14:1141–1152. doi: 10.1523/JNEUROSCI.14-03-01141.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Ohtoshi A, Wang SW, Maeda H, Saszik SM, Frishman LJ, et al. Regulation of retinal cone bipolar cell differentiation and photopic vision by the CVC homeobox gene Vsx1. Curr Biol. 2004;14:530–536. doi: 10.1016/j.cub.2004.02.027. [DOI] [PubMed] [Google Scholar]
  76. Ornitz DM, Marie PJ. FGF signaling pathways in endochondral and intramembranous bone development and human genetic disease. Genes Dev. 2002;16:1446–1465. doi: 10.1101/gad.990702. [DOI] [PubMed] [Google Scholar]
  77. Pan Q, Bakowski MA, Morris Q, Zhang W, Frey BJ, et al. Alternative splicing of conserved exons is frequently species-specific in human and mouse. Trends Genet. 2005;21:73–77. doi: 10.1016/j.tig.2004.12.004. [DOI] [PubMed] [Google Scholar]
  78. Pin JP, Waeber C, Prezeau L, Bockaert J, Heinemann SF. Alternative splicing generates metabotropic glutamate receptors inducing different patterns of calcium release in Xenopus oocytes. Proc Natl Acad Sci USA. 1992;89:10331–10335. doi: 10.1073/pnas.89.21.10331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Perrin RM, Konopatskaya O, Qiu Y, Harper S, Bates DO, Churchill AJ. Diabetic retinopathy is associated with a switch in splicing from anti- to pro-angiogenic isoforms of vascular endothelial growth factor. Diabetologia. 2005;48:2422–2427. doi: 10.1007/s00125-005-1951-8. [DOI] [PubMed] [Google Scholar]
  80. Pyper JM, Bolen JB. Identification of a novel neuronal C-SRC exon expressed in human brain. Mol Cell Biol. 1990;10:2035–2040. doi: 10.1128/mcb.10.5.2035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Rauen T, Wiessner M, Sullivan R, Lee A, Pow DV. A new GLT1 splice variant: cloning and immunolocalization of GLT1c in the mammalian retina and brain. Neurochem Int. 2004;45:1095–1106. doi: 10.1016/j.neuint.2004.04.006. [DOI] [PubMed] [Google Scholar]
  82. Ravindranathan A, Parks TN, Rao MS. New isoforms of the chick glutamate receptor subunit GluR4: molecular cloning, regional expression and developmental analysis. Brain Res Mol Brain Res. 1997;50:143–153. doi: 10.1016/s0169-328x(97)00179-4. [DOI] [PubMed] [Google Scholar]
  83. Resch A, Xing Y, Alekseyenko A, Modrek B, Lee C. Evidence for a subpopulation of conserved alternative splicing events under selection pressure for protein reading frame preservation. Nucleic Acids Res. 2004;32:1261–1269. doi: 10.1093/nar/gkh284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Rice DS, Nusinowitz S, Azimi AM, Martinez A, Soriano E, Curran T. The reelin pathway modulates the structure and function of retinal synaptic circuitry. Neuron. 2001;31:929–941. doi: 10.1016/s0896-6273(01)00436-6. [DOI] [PubMed] [Google Scholar]
  85. Ruegg MA, Tsim KWK, Horton SE, Kroger S, Escher G, et al. The agrin gene codes for a family of basal lamina proteins that differ in function and distribution. Neuron. 1992;8:691–699. doi: 10.1016/0896-6273(92)90090-z. [DOI] [PubMed] [Google Scholar]
  86. Rupp F, Ozcelik T, Linial M, Peterson K, Francke U, Scheller R. Structure and chromosomal localization of the mammalian agrin gene. J Neurosci. 1992;12:3535–3544. doi: 10.1523/JNEUROSCI.12-09-03535.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Sato M, Kitazawa T, Katsumata A, Mukamoto M, Okada T, Takeya T. Tissue-specific expression of two isoforms of chicken fibroblast growth factor receptor, bek and Cek3. Cell Growth Differ. 1992;3:355–361. [PubMed] [Google Scholar]
  88. Schmitt A, Asan E, Lesch KP, Kugler P. A splice variant of glutamate transporter GLT1/EAAT2 expressed in neurons: cloning and localization in rat nervous system. Neuroscience. 2002;109:45–61. doi: 10.1016/s0306-4522(01)00451-1. [DOI] [PubMed] [Google Scholar]
  89. Schuur ER, Dasgupta P, Reddy EP, Rabinovich JM, Baluda MA. Alternative splicing of the chicken c-myb exon 9A. Oncogene. 1993;8:1839–1847. [PubMed] [Google Scholar]
  90. Schwarz DA, Barry G, Eliasof SD, Petroski RE, Conlon PJ, Maki RA. Characterization of gamma-aminobutyric acid receptor GABAB(1e), a GABAB(1) splice variant encoding a truncated receptor. J Biol Chem. 2000;275:32174–32181. doi: 10.1074/jbc.M005333200. [DOI] [PubMed] [Google Scholar]
  91. Stohr H, Berger C, Frohlich S, Weber BH. A novel gene encoding a putative transmembrane protein with two extracellular CUB domains and a low-density lipoprotein class A module: isolation of alternatively spliced isoforms in retina and brain. Gene. 2002;286:223–231. doi: 10.1016/s0378-1119(02)00438-9. [DOI] [PubMed] [Google Scholar]
  92. Thanaraj TA, Clark F, Muilu J. Conservation of human alternative splice events in mouse. Nucleic Acids Res. 2003;31:2544–2552. doi: 10.1093/nar/gkg355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Thomas WS, O’Dowd DK, Smith MA. Developmental expression and alternative splicing of chick agrin RNA. Dev Biol. 1993;158:523–535. doi: 10.1006/dbio.1993.1210. [DOI] [PubMed] [Google Scholar]
  94. Toda M, Shirao T, Minoshima S, Shimizu N, Toya S, Uyemura K. Molecular cloning of cDNA encoding human drebrin E and chromosomal mapping of its gene. Biochem Biophys Res Commun. 1993;196:468–472. doi: 10.1006/bbrc.1993.2273. [DOI] [PubMed] [Google Scholar]
  95. Tominaga S, Kuroiwa K, Tago K, Iwahana H, Yanagisawa K, Komatsu N. Presence and expression of a novel variant form of ST2 gene product in human leukemic cell line UT-7/GM. Biochem Biophys Res Commun. 1999;264:14–18. doi: 10.1006/bbrc.1999.1469. [DOI] [PubMed] [Google Scholar]
  96. Trabucchi M, Tostivint H, Lihrmann I, Blahser S, Vallarino M, Vaudry H. Characterization of the cDNA encoding a somatostatin variant in the chicken brain: comparison of the distribution of the two somatostatin precursor mRNAs. J Comp Neurol. 2003;461:441–451. doi: 10.1002/cne.10690. [DOI] [PubMed] [Google Scholar]
  97. Turner DL, Cepko CL. A common progenitor for neurons and glia persists in rat retina late in development. Nature. 1987;328:131–136. doi: 10.1038/328131a0. [DOI] [PubMed] [Google Scholar]
  98. Ulaner GA, Hu JF, Vu TH, Giudice LC, Hoffman AR. Telomerase activity in human development is regulated by human telomerase reverse transcriptase (hTERT) transcription and by alternate splicing of hTERT transcripts. Cancer Res. 1998;58:4168–4172. [PubMed] [Google Scholar]
  99. Wagner N, Wagner KD, Hammes A, Kirschner KM, Vidal VP, et al. A splice variant of the Wilms’ tumour suppressor Wt1 is required for normal development of the olfactory system. Development. 2005;132:1327–1336. doi: 10.1242/dev.01682. [DOI] [PubMed] [Google Scholar]
  100. Wetts R, Fraser SE. Multipotent precursors can give rise to all major cell types of the frog retina. Science. 1988;239:1142–1145. doi: 10.1126/science.2449732. [DOI] [PubMed] [Google Scholar]
  101. Wistow G, Bernstein SL, Wyatt MK, Ray S, Behal A, et al. Expressed sequence tag analysis of human retina for the NEIBank Project: retbindin, an abundant, novel retinal cDNA and alternative splicing of other retina-preferred gene transcripts. Mol Vis. 2002;8:196–204. [PubMed] [Google Scholar]
  102. Woo CH, Sopchak L, Lipsick JS. Overexpression of an alternatively spliced form ofc-Myb results in increases in transactivation and transforms avian myelomonoblasts. J Virol. 1998;72:6813–6821. doi: 10.1128/jvi.72.8.6813-6821.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Woolard J, Wang WY, Bevan HS, Qiu Y, Morbidelli L, Pritchard-Jones RO, et al. VEGF165b, an inhibitory vascular endothelial growth factor splice variant: mechanism of action, in vivo effect on angiogenesis and endogenous protein expression. Cancer Res. 2004;64:7822–7835. doi: 10.1158/0008-5472.CAN-04-0934. [DOI] [PubMed] [Google Scholar]
  104. Yong TJ, Gan YY, Toh BH, Sentry JW. Human CKIalpha(L) and CKIalpha(S) are encoded by both 2.4- and 4. 2-kb transcripts, the longer containing multiple RNA-destabilizing elements. Biochim Biophys Acta. 2000;1492:425–433. doi: 10.1016/s0167-4781(00)00146-9. [DOI] [PubMed] [Google Scholar]
  105. Yoshida H, Goedert M. Molecular cloning and functional characterization of chicken brain tau: isoforms with up to five tandem repeats. Biochemistry. 2002;41:15203–15211. doi: 10.1021/bi026464m. [DOI] [PubMed] [Google Scholar]
  106. Zhang J, Gross SD, Schroeder MD, Anderson RA. Casein kinase I alpha and alpha L. alternative splicing-generated kinases exhibit different catalytic properties. Biochemistry. 1996;35:16319–16327. doi: 10.1021/bi9614444. [DOI] [PubMed] [Google Scholar]
  107. Zhao X, Huang J, Khani SC, Palczewski K. Molecular forms of human rhodopsin kinase (GRK1) J Biol Chem. 1998;273:5124–5131. doi: 10.1074/jbc.273.9.5124. [DOI] [PubMed] [Google Scholar]
  108. Zhao X, Yokoyama K, Whitten ME, Huang J, Gelb MH, Palczewski K. A novel form of rhodopsin kinase from chicken retina and pineal gland. FEBS Lett. 1999;454:115–121. doi: 10.1016/s0014-5793(99)00764-4. [DOI] [PubMed] [Google Scholar]

RESOURCES