Split Gene
Split Gene
INTRODUCTION
By the late 70s the physical structure of a gene was firmly established from
work in bacteria. The sequences of the gene, the RNA and the protein were
colinearly organized and expressed. Since the science of genetics suggested
that genes in eukaryotic organisms behaved similarly to those of prokaryotic
organisms, it was naturally assumed that this bacterial gene structure was
universal. It followed that if the gene structure was the same, then the
mechanisms of regulation were probably very similar, and thus what was
true of a bacterium would be true of an elephant.
However, many descriptive biochemical aspects of the genetic material
and its expression in cells with nuclei suggested that the simple molecular
biology of gene expression in bacteria might not be universal. First, obviously, RNAs transcribed from nuclear genes are physically separated from
the translational machinery in the cytoplasm. Thus, the nuclear compartment could be the site of selective RNA processing and transport. Second,
the DNA content of eukaryotic germ cells varied significantly between
organisms without an apparent variation in the number of genes. Some
organisms appeared to have ten times as much DNA as was required to
encode all of the proteins. Third, the previously described phenomenon of
heterogeneous nuclear RNA (hnRNA) suggested that long RNAs were
transcribed from diverse nuclear sequences (1). These hnRNAs had a short
half-life relative to cytoplasmic mRNAs and thus could potentially be precursors to mRNAs. Furthermore, both the long hnRNA and the shorter
mRNA appeared to have modified 5 and 3 termini in common, a
7mGpppX cap (2, 3, 4) and polyadenylation tracts (5, 6, 7), respectively. The
meaning of these observations concerning hnRNAs remained controversial
as it was not possible to establish a precursor-product relationship between
the nuclear RNA population and the cytoplasmic mRNA.
Whether or not the structure of genes in eukaryotic cells was the same as
that in bacteria was not really questioned at that time. The important issue
was to establish the exact biochemical pathway between a gene in the
nucleus and its mRNA in the cytoplasm. This hypothetical pathway was
146
A
I
I
0
GC
I
I
10
B
I
I
I
I
20
I
30
J D
II
I
I
40
I
50
F D E
A
I
I
60
I
70
I Eco RI
H L E
F K
II
I
I I Hind III
I
80
I
90
I MAP UNIT
100
35,000 bp of adcnovirus 2 DNA was assigned as 100 map units. The positions of cleavage sites
are denoted by vertical lines and fragments are lettered on the basis of length. The r and 1 viral
strands are transcribed to the right and left, respectively. The sequences constituting the body of
mRNA specifying the abundant hexon (II) protein are encompassed by the bar about the
boundary of EcoRI fragments A and B.
Phillip A. Sharp
the latterR-toop hybrids observed after incubation of hexon mRNA and duplex HindIII A
fragment DNA arc shown in A and B and is diagrammed schematically in C. Similarly, two
examples of hybrids of hexon mRNA and the single-stranded HindIII A fragment are shown in
D and E. A schematic. of the hybrid structure shown in E is given in F. The single-stranded RNA
at the end of the hybrid region is represented by a wave-like line. In A, B, D, and E the positions
of the KNA tails at the 5 and 3 ends of the hybrids are denoted by arrows. An example of a
hybrid between single-stranded EcoRI A DNA and hexon RNA is shown in G and diagrammed in
H. The hybrid region is indicated by a heavy line; loops A, B, and C (single-stranded unhybridized DNA) are joined by hybrid regions resulting from annealing of upstream DNA sequences
to the 5 tail of hexon mRNA. Bars on micrographs represent 0.1 M.
148
RNA was transcribed from the adenovirus genome during the late stages of
infection (12) and the stable cytoplasmic RNAs were shorter than the
predominant nuclear RNAs. Thus, the production of RNAs during the late
stage of adenovirus infection presented a paradigm for the heterogeneous
nuclear RNA phenomenon associated with cellular genes.
DISCOVERY OF RNA SPLICING
Comparison of the structure of a cytoplasmic mRNA to that of a nuclear
precursor RNA (13) required the purification of a specific homogeneous
mRNA. The most abundant mRNA, which encoded the adenovirus hexon
protein, was separated from other viral mRNAs by gel electrophoresis and
used in electron microscope mapping studies (14). Ray White, as a fellow at
Stanford, was the first to recognize that RNA.DNA hybrids were more
stable than DNA.DNA duplexes in high concentrations of formamide (15).
This phenomenon was characterized physically by Davidsons lab (16) and
was the basis of a convenient R-looping technique whereby RNA forms a
hybrid with a DNA strand, displacing the other strand of DNA into an easily
observed loop. The adenovirus mRNA for hexon protein was mapped by
the R-loop method to the HindIII A fragment of adenovirus 2 (13).
Inspection of the R-loops between the hexon-mRNA and the HindIII B
DNA fragment revealed the presence of RNA tails at both the 5 and 3 ends
of the hybrid (Fig. 2A, B and C). The single strand RNA tail at the 3 end was
expected, as the RNA was known to be polyadenylated post-transcriptionally. The single strand RNA tail at the 5 end of the hybrid was not expected;
however, this 5 tail of RNA could have been displaced from the RNA.DNA
hybrids by formation of duplex DNA by a process called branch migration.
In fact, similar 5 tails had been observed previously in R-loop mapping of
adenovirus mRNA and had been ascribed to such branch migration (17,
18). Arguing against branch migration however was the finding that the
lengths of the 5 tails were relatively uniform, 170 nucleotides. To eliminate
this potential of competing DNA sequences displacing RNA sequences, we
used a denatured single-strand of the HindIII fragment (Fig. 2D, E, and F)
to form a RNA.DNA hybrid. Surprisingly, the 5 RNA tail still did not form
a hybrid with the adjacent viral sequences, suggesting that these RNA
sequences were derived from other DNA sequences.
Could the DNA sequences transcribed to form the 5 tail sequence, the
leader RNA sequence, be located upstream of the body of the hexon
mRNA, perhaps as part of the long nuclear RNA? To test this possibility, a
strand of the EcoRI A DNA fragment was hybridized with the hexon mRNA.
This fragment contained all of the linear viral sequences which could have
been transcribed by the polymerase before encountering the body of the
hexon mRNA. Surprisingly, and wonderfully, the leader sequences hybridized to three short tracts of DNA sequences, creating three different size
loops A, B and C of intervening single-strand DNA (Fig. 2G and H). The
length of these loops mapped the positions of the leader sequences, Ll , L2,
149
Richard J. Roberts
L3, to approximately 16.9, 19.8, and 26.9 map units on the genome,
respectively. The distances between the bases of the three loops permitted
an estimate of the lengths of the two internal leaders to be 80 and 110
nucleotides, respectively. The 5 proximal leader was estimated to be quite
short, but longer than fifteen nucleotides.
RNA splicing was the mechanism we proposed for generation of the final
hexon mRNA (13). It was known that the nucleus of virus-infected cells
contained long RNAs transcribed from the viral sequences between the 5
most leader Ll , located at 17 map units, and the body of the hexon mRNA,
located between 51.7 and 61.3 map units (19, 20). Thus, the long nuclear
RNA probably contained sequences for all three leaders, Ll , L2 and L3, as
well as for the body of the mRNA. These sequences were conjectured to be
joined by excision of the intervening sequences and ligation of the flanking
RNAs, a process dubbed RNA splicing (Fig. 3). In fact, the nuclear RNAs
were quite abundant and were easily visualized by electron microscopy of
the RNA.DNA hybrids (21). Analysis of the structure of these RNA.DNA
hybrids revealed the presence of potential intermediates, where only Ll had
been spliced to L2 and where only Ll, L2 and L3 had been joined by
splicing:
The RNA splicing hypothesis reconciled many paradoxes. Both the longer nuclear RNAs and the shorter cytoplasmic mRNAs could share 5 cap
termini, 7 mG pppX, and 3 po1yA tracts (22), because internal sequences
were removed from the nuclear precursor. More specifically for adeno-
17 20
27
m
u
52
c
hexon
(AIn
hexon
(A)n
L123
hexon
(A)n
(A)n
mature hexon mRNA
Fig. 3: Proposed RNA splicing mechanism for synthesis of mRNA for hexon protein. A long nuclear
precursor RNA is transcribed from 16.9 map units through the polyA addition site at the end of
the body of hexon mRNA. The region of the Ad2 genome from which the precursor RNA is
transcribed is shown at the top of the figure. The four RNA segments in the cytoplasmic mRNA
are processed from this precursor by excision of intervening sequences (denoted by dashed arrows).
150
virus, previous evidence had suggested that many different viral mRNAs
appeared to share a common 5 terminal sequence (23). This sequence, a
long Tl oligonucleotide, contained the cap and an apparently unique
sequence eleven residues in length. If many of the late viral mRNAs were
processed by splicing from the same type of precursor RNA, then they
could share a common sequence at their 5 termini. More importantly, the
RNA splicing hypothesis provided an explanation for the hnRNA phenomenology associated with cellular genes. Heterogeneous nuclear RNA transcribed from diverse cellular genes could be processed by RNA splicing into
shorter cytoplasmic mRNAs. Thus, most cellular genes probably contained
sequences which were removed by RNA splicing, i.e. they were split genes.
SPLIT GENES: INTERVENING SEQUENCES OR INTRONS IN
CELLULAR GENES
Shortly after the discovery of RNA splicing and split genes in adenovirus, a
number of cellular genes were also shown to have introns or intervening
sequences. For example, the globin genes contained two intervening sequences (24, 25), the ovalbumin gene was split into eight sets of sequences
(26), and the immunoglobulin genes contained both short and long introns
(27). In fact, the average cellular gene contains approximately eight introns
and the primary transcription unit is typically four times larger than the
final mRNA. Shortly thereafter, it was recognized that there was a limited
set of conserved sequences at each intron boundary (28). Interestingly,
these consensus sequences were common of vertebrate, plant and yeast cells
(29) suggesting the splicing process was evolutionarily general. Introns in
the latter organisms are generally shorter and have more highly conserved
sequences at their boundaries.
Phylogenetic comparison of the sequences of homologous genes from a
variety of organisms revealed that intron sequences had drifted much more
rapidly than exon sequences. This suggested that intron sequences are
generally not functional, at least in the context of requiring long tracts of
specific sequences. Furthermore, the length of introns in homologous
genes varied significantly during evolution, suggesting little constraint.
Finally, it was clear that specific introns could be lost during evolution. The
mechanism responsible for the exact deletion of introns is probably related
to gene conversion using a cDNA copy of the mRNA or a partially spliced
intermediate RNA. This process has been documented for the removal of
introns from yeast genes (30) and raises the question of why introns persisted during evolution.
MUTATIONS IN CELLULAR GENES
Many human diseases are caused by mutations that interfere with RNA
splicing. Approximately one quarter of all mutations in the human globin
genes underlying thalassemia are mutations in sequences specifying correct
151
Phillip A. Sharp
cryptic
5 splice sites
-t
new exon
Fig. 4: -Globin mutations. -Globin genes of thalassemia patients possessing mutations which
affect RNA splicing (31). The single base changes in four independent patients are indicated by
arrows from the diagram of the two intron structure of the -globin gene. Three of the
mutations alter conserved sequences of the 5 splice site and a fraction of the mRNA from these
mutant genes are processed at the cryptic 5 splice sites. The fourth mutation is within the
sequences of the second intron and creates a 5 splice site at this position. This results in the
induction of the indicated sequences as a new exon. Exon sequences are represented by
rectangles, intron sequences by lines.
152
Fibronectin Protein:
Fibronectin Gene:
Protein Subdomain
Type I
Type ll ()
Type Ill 0
Fig. 5: Fibronectin gene evolved by exon duplication. The large extracellular matrix protein, fibronectin, is primarily composed of three types of protein domains, I, II, and III denoted as tilted
ellipsoids, vertical ellipsoids and circles, respectively. As indicated at the bottom, homologous
domains are found in other cellular proteins. Each of these domains are encoded by a defined
exon pattern, one or two exons, denoted as verticle rectangles in the middle. These exon units
are duplicated in the ftbronectin gene and also in the other genes containing these domains. In
two cases, larger rectangles, the intron separating the typical two exon structures of the type III
subdomain, have been deleted. These patterns suggest that descendants of a common progenitor exon configuration was used to generate parts of these genes (36).
Phillip A. Sharp
153
FIBROBLASTS
LIVER
Fig. 6: Alternative splicing of RNA encoding fibronectin. The exons denoted as EIIIB and EIIIA are
spliced into the mature RNA synthesized in many cell types including fibroblasts. However, in
the liver, these two exons are not included, are skipped, in synthesis of the mRNA. Furthermore, five variations for the splicing patterns of the V region are shown. All of these variations
on splicing are found in most cell types. The fibronectin proteins encoded by the first and third
mRNA bind differentially to receptors on lymphocytes. In total, at least twenty different
proteins are synthesized from the single fibronectin gene.
154
Phillip A. Sharp
Fig. 7: The structure of pyruvate kinase and intron positions. Pyruvate kinase is shown schematically
as a protein with secondary and tertiary structure on the left. On the right, this structure is
expanded at the positions of introns in the tertiary structure of the protein. Note that the first
three introns all fall between a-helix (barrels) and -sheet (arrows) repeats. Intron 8 is positioned in approximately equivalent positions in the mononucleotide binding fold of PK and
maize alcohol dehydrogenase. Reprinte d with permission from ref 44.
both prokaryotic and eukaryotic organisms may have had a split gene
structure more typical of a current eukaryotic cell (for discussion see 44).
IS THE GENE AN EXON?
The gene was first genetically defined as a unit of inheritance which is
associated with a locus on a chromosome. The chemical definition of a gene
has become much more difficult however, as the complexity of genetic
information and its modifications are discovered. The existence of alternative splicing of exons, where information at an exon unit can be optionally
expressed in some cells and not in others, suggests that an exon might
correspond to a gene. That is, an exon corresponds to the minimal amount
of information which is expressed as a discrete unit. This concept becomes
particularly relevant when the trans-splicing process is considered (45). In
this case, exons transcribed from different loci, and in many cases different
chromosomes, are joined by RNA splicing. Trans-splicing of exons and
introns has been established in the parasite trypanosomes (46), the flat
worm C. elegans (47), and suggested for some human genes (48). Clearly, in
the case of trans-splicing, both the unit of inheritance and the locus on a
chromosome can correspond to a single exon.
156
Phillip A. Sharp
Spliceosome
Nuclear pre-mRNA
-p-
157
Self-splicing
Group
P-
-Ip-
Fig. 8: Comparison of self-splicing and nuclear pre-mRNA splicing mechanism. The first column
outlines the mRNA precursor splicing mechanism. The shaded circle represents a multicomponent complex, the spliceosome, which promotes the splicing reaction. The second column
outlines the splicing mechanism of self-splicing introns of the group I type. This process is
catalyzed by RNA structures within the intron (dark semicircle), which contains a guanosine
binding site, and utilizes a guanosine (C) factor in the first step. The third column outlines the
splicing mechanism of self-splicing introns of the group II type. This process is also catalyzed by
RNA structures within the intron but utilizes, instead of a cofactor, an adenosine residue (A)
within the intron to form a lariat RNA. All three mechanisms proceed by two steps, reaction at
the 5 splice site and then reaction at the 3 splice site. The fate of the phosphates at the 5 and 3
splice sites is indicated.
158
Phillip A. Sharp
Fig. 9: RNA interactions between spliceosomal snRNAs and pre-mRNA substrates. (A) (Top) Basepairing interactions between Ul and U2 snRNAs and pre-mRNA are indicated on left and right
of intron, respectively. (Bottom) Extensive base-pairing between U4 and U6 snRNAs. (B) Interactions between Ul, U2, U5, and U6 snRNAs and pre-mRNA. In both A and B, pre-mRNA
consensus sequences and snRNA sequences are those of S. cerevisiae;
uppercase nucleotides are
highly conserved between S. cerevisiae
and the known sequences of other organisms (excluding
trypanosomes, which do not have a GUAGUA sequence in U2 snRNA). Different shaded areas
highlight sequences in U2 and U6 snRNAs that change base-pairing partners during the
spliceosome cycle. Internal snRNA secondary structures that do not change between A and B
are shown as stylized stems and loops. Asterisks indicate snRNA positions at which mutations
specifically block the second step of splicing.
160
Phillip A. Sharp
5 ss
bp
3 ss
Py
a BAG
5--GU
B - 3
Pre-mRNA
5_ - . ..--q-3
mRNA
Fig. 10: Schematic representation of the spliceosome cycle in terms of the role of small nuclear ribonucleoprotein (snRNP) particles in pre-mRNA splicing. Pre-mRNA (top line), containing two exons
separated by an intron, enters splicing complexes with snRNPs and exits as mRNA (bottom line)
and excised lariat intron (left border ). Other, non-snRNP factors are required for spliceosome
formation, but have been omitted for simplicity. CC, A, Bl, B2, Cl, C2, and I represent
complexes within the splicing pathway that have been distinguished biochemically and/or
genetically. 5 SS, 3 SS, bs, and Py indicate 5 and 3 splice sites, branch site, and polypyrimidine
tract, respectively. The individual snRNPs indicated are Ul, U2, U4, U5, and U6.
162
ss
5--GU
bp
3 ss
PY
A BAG--~
Pre-mRNA
mRNA
Fig. 11: Transitions in the spliceosome cycle which require a PRPprotein. The particular PRP mutant
is listed beside the arrow indicating the transition in the cycle in vitro which requires the mutant
protein.
Phillip A. Sharp
163
recycled for further splicing in vivo, while the intron RNA is degraded.
Since all snRNAs are very stable, a particular snRNA likely participates in
many splicing cycles.
Assembly and rearrangements of complexes as elaborate as those in the
spliceosome cycle would be expected to require a number of proteins.
Sophisticated genetic analysis has identified a number of yeast gene products that are important for splicing either in vivo or in vitro (82, 83).
Temperature sensitive mutations in PRP (precursor RNA processing) genes
cause the accumulation of unspliced pre-mRNA in the nucleus and/or
generate extracts defective for splicing in vitro (Fig. 11). Surprisingly, it is
estimated that at least a hundred different genes encode products important for RNA splicing, about 2% of the total yeast genome. Thus, the
splicing apparatus is a significant component of the nucleus. The amino
acid sequences predicted for some of the PRP genes suggests that they have
functions such as RNA binding, RNA helicase and protein-protein interaction properties. Matching these hypothetical functions to processes in spliceosomes will be challenging.
Reassuringly, the steps in the spliceosome cycle where particular PRP
proteins are required (Fig. 11) are consistent with the cycle as defined by
kinetic and biochemical methods. Most transitions between specific forms
of the spliceosome require one or more specific proteins. Furthermore, a
number of PRP mutants are defective in splicing because of their inability to
reassemble snRNPs for further splicing. Thus, both genetic and biochemical results prove that the spliceosome cycle is the process responsible for
excision of introns from split genes.
THE CHEMISTRY OF THE SPLICEOSOME
Both steps in the spliceosome process involve reactions between hydroxyl
groups and a phosphodiester bond. The products and substrates in both
steps contain the same number of covalent bonds and thus each step could
be accomplished with a single transesterification reaction. This has been
shown to be the case for group I self-splicing introns by analysis of the
stereochemistry of the two reactions (Fig. 12; 84, 85, 86). Phosphorothioates in diester bonds of RNA are chiral, having either a Rp or Sp
stereochemistry depending upon the position of the sulfur atom. The
specific stereochemistry can be determined by its sensitivity to particular
nucleases. If a chiral phosphorothioate group participates in a single transesterification reaction, its stereochemistry is inverted, for example from
Rp to Sp. This fact permits counting of the number of transesterification
reactions in a process and, thus, the detection of any potential transient
intermediate.
As implied above, Rp and Sp phosphorothioates are not equally active in
many catalytic sites. This is commonly interpreted as reflecting the unique
role of one of the oxygen groups on the phosphate in interaction with a
metal ion, a role which could not be equally well-fulfilled by a sulfur group.
164
-Binding Site
G-lvs
-Binding Site
Fig. 12: The stereochemistry of the first and second step reactions of the group I self-splicing site. A sulfur
atom is indicated at the positions where it does not inhibit the reactions, Rp and Sp positions for
the first and second steps, respectively (84, 85, 86). The interactions between the oxygen group
on the phosphate and the Mg+ + groups are hypothetical, proposed to explain why a phosphorothioate with the sulfur atom in this position is not reactive.
165
Phillip A. Sharp
Group I
Spliceosome
RP
tl
sp
1st step
t1
1st step
t1
2nd step
@I
G
P-
y?H
SP
t1
RP
2nd step
Gp&
-p-
Fig. 13: Chemical mechanisms and stereochemical configurations of the two steps of pre-mRNA splicing
(right) and group I intron self-splicing (left) First-step differences between spliceosomal and group
166
,speckles
poly A+ RNA
Ul. U2. LE. U4/U 6snRNPs
SC35. SF2IASF
Ul snRNP
U2AF
coiled bodies
Ul. U2. U5. U4/U 6snRNPs
U2AF
Fig. 14: The nucleus. Subnuclear localization of splicing factors. Specific antisera and in situ
hybridization methods have been used to localize components of the spliceosome and splicing
factors in the nucleus of mammalian cells. The outline of the cell is shown without structure in
the cytoplasm. Three nuclear pores are diagrammed in the nuclear membrane. The large dark
structures are nucleoli,.site of ribosomal RNA synthesis. Some of the components of the
speckles, regions containing pre-mRNA, are listed. The coiled bodies are different and do not
appear to contain pre-mRNAs (93). The general shadowing represents the nuclear distribution
of Ul snRNP and U2AF.
14; 91, 92). This contrasts with the subnuclear location of Ul snRNP,
which, although also concentrated in speckles and foci, is more uniformly
dispersed throughout the nucleus. These speckled structures correspond to
regions described by electron microscopists as interchromatin granule clusters and the perichromatin fibril network. High resolution in situ hybridization methods and pulse labeling studies locate newly synthesized RNA in the
speckled regions and also in curvilinear tracks which extend from the gene
toward the nuclear periphery (93, 94). The further the nuclear precursor
RNA is located along the tracks which lead from its gene of origin, the lower
the relative concentration of intron sequences as compared to exon sequences. This suggests that the precursor RNA moves through the speckled
structures as it is being processed and transported (94). Active spliceosomes
are probably concentrated in these regions.
Several proteins important for RNA splicing are also concentrated in
speckled structures with the snRNPs. These include the Arg/Ser proteins
SF2/ASF (95, 96) and SC35 (97). Surprisingly, these proteins, as well as the
snRNPs in active spliceosomes, are probably attached to an operationally
defined nuclear matrix. This matrix is the structure that remains in the
168
nucleus after almost all of the chromatin proteins and DNA are extracted by
sequential treatments with DNase 1 and high salt (98). The integrity of the
fibrillar matrix which remains, is sensitive to RNase suggesting a major role
for RNA in its structure. Spliceosomes have been shown to be associated
with the nuclear matrix as it has been possible to process pulse-labeled
endogenous precursor RNA extracted as part of the matrix (98).
Active spliceosomes may be associated with the nuclear matrix through
interactions with proteins in the Arg/Ser family (99, 100). A panel of
monoclonal antibodies has been generated to components in the nuclear
matrix of human cells. Three of these monoclonal antibodies specifically
stain the matrix and extensively co-localize with the snRNPs in the speckled
structures. Surprisingly, these three antibodies specifically immunoprecipitate spliceosome complexes which contain exon sequences (101). Thus, the
antibodies immunoprecipitate a fraction of the precursor RNA, the lariat
intermediate RNA and the associated 5 exon, and the spliced exon product
RNAs. The antibodies do not immunoprecipitate the intron containing
spliceosome complex I even though it contains most of the snRNPs and
associated protein components. These remarkable results suggest that the
matrix components bind to exon RNA sequences that are specifically assembled into splicing complexes.
The proteins recognized by the monoclonal antibodies to the nuclear
matrix are most likely members of the Arg/Ser family of proteins or are
associated with the Arg/Ser proteins. Precursor RNAs are not immunoprecipitated from reactions containing nuclear extracts depleted of Arg/Ser
proteins. Further addition of preparations of purified Arg/Ser proteins will
block specific immunofluoresence staining of subnuclear structures. These
results are consistent with a model where the Arg/Ser family of proteins are
associated with or form the nuclear matrix (101). Newly synthesized RNA
sequences are recognized by the Arg/Ser proteins such as SC35, ASF, the
U1 70 kd and UPAF and become associated with the matrix, perhaps,
through interactions of the Arg/Ser tracts. On this structure, complete
spliceosomes would form encompassing the matrix proximal splice sites and
execute splicing. The exon product RNAs would remain matrix bound and
move by some unknown mechanism to the nuclear pore.
CONCLUSION
The discovery of split genes and RNA splicing has been critical for studies of
the biology of eukaryotic organisms. Gene regulation is central to all biological phenomena and RNA splicing is important in the regulation of
genes, particularly when precursor RNAs are processed by alternative pathways to generate mRNAs encoding different proteins. The mechanism of
splicing by the spliceosome is probably related to the self-splicing process of
group I and II introns. The spliceosome process is old in an evolutionary
sense, perhaps as old as the ribosomal process responsible for translation.
Thus, the eukaryotic cell can be conjectured as consisting of two compart-
Phillip A. Sharp
169
REFERENCES
1.
2.
2069-2071.
18. Chow, L. T., Roberts, J. M., Lewis, J. B. and Broker, T. R. (1977) Cell, 11,
819-836.
19. Goldberg, S., Weber, J. and Darnell, J. E., Jr. (1977) Cell, 10, 617-622.
20. Weber, J., Jelinek, W. and Darnell, J. E., Jr. (1977) Cell, 10, 611-616.
21. Berget, S. M. and Sharp, P. A. (1979)J. Mol. Biol., 129, 547-565.
22. Perry, R. P. and Kelley, D. E. (1976) Cell, 8, 433 -442.
23. Gelinas, R. E. and Roberts, R. J. (1977) Cell, 11, 533-544.
24. Jeffreys, A. J. and Flavell, R. A. (1977) Cell, 12, 1097- 1108.
25. Tilghman, S. M., Tiemeier, D. C., Seidman, J. G., Peterlin, B. M., Sullivan, M.,
Maizel, J. V. and Leder, P. (1978) Proc. Natl. Acad. Sci. USA, 75, 725 - 729.
26. Breathnach, R., Manel, J-L. and Chambon, P. (1977): Nature, 270, 314-319.
27. Tonegawa, S., Maxam, A. M., Tizard, R., Bernard, 0. and Gilbert, W. (1978)
Proc. Natl. Acad. Sci. USA, 75, 1485 - 1489.
28. Breathnach, R. and Chambon, P. (1981) Annu. Rev. Biochem., 50, 349-384.
29. Padgett, R. A., Grabowski, P. J., Konarska, M. M., Seiler, S. and Sharp, P. A.
(1986) Annu . Rev. Biochem., 5 5, 1119-1150.
170
Phillip A. Sharp
171
68. Padgett, K. A., Mount, S. M., Steitz, J. A. and Sharp, P. A. (1983) Cell, 35,
101 - 107.
69. Parker, R., Siliciano, P. G. and Guthrie, C. (1987) Cell, 49, 229-239.
70. Guthrie, C. and Patterson, B. (1988) Annu. Rev. Genet., 22, 387-419.
71. Konarska, M. M. and Sharp, P. A. (1987) Cell, 49, 763-774.
72. Madhani, H. D. and Guthrie, C. (1992) Cell, 71, 803-817.
73. McPheeters, D. S. and Abelson, J. (1992) Cell, 71, 819-831.
74. Sawa, H. and Abelson, J. (1992) Proc. Natl. Acad. Sci. USA, 89, 11269- 11273.
75. Wassarman, D. A. and Steitz, J. A. (1992) Science, 257, 1918- 1925.
76. Newman, A. and Norman, C. (1992) Cell, 68, 743-754.
77. Moore, M. J., Query, C. C. and Sharp, P. A. (1993) In The RNA World, Cold
Spring Harbor Laboratory Press, pp. 303 - 357.
78. Seraphin, B. and Rosbash, M. (1991) EMBO J., 10, 1209-1216.
79. Michaud, S. and Reed, R. (1991). Genes Dev., 5, 2534-2546.
80. Jamison, S. F. and Garcia-Blanco, M. A. (1992) Proc. Natl. Acad. Sci. USA, 89,
5482 - 5486.
81. Ruskin, B., Zamore, P. D. and Green, M. R. (1988) Cell, 52, 207-219.
82. Ruby, S. W. and Abelson, J. (1988) Trends Genet., 7, 79-85.
83. Guthrie, C. (1991) Science, 253, 157-163.
84. McSwiggen, J. A. and Cech, T. R. (1989) Science, 244, 679-683.
85. Rajagopal, J., Doudna, J. A. and Szostak, J. W. (1989) Science, 244, 692-694.
86. Suh, E.-R. and Waring, R. B. (1992) Nucleic Acids Res., 20, 6303 -6309.
87. Moore, M. J. and Sharp, P. A. (1993) Nature, 365, 364- 368.
88. Moore, M. J. and Sharp, P. A. (1992) Science, 256, 992-997.
89. Beyer, A. L. and Osheim, Y. N. (1988) Genes Dev., 2, 754- 765.
90. Mehlin, H., Daneholt, B. and Skoglund, U. (1992) Cell, 69, 605-613.
91. Spector, D. L. (1993) Annu. Rev. Cell Biol., 9, 265-315.
92. Nyman, U., Hallman, I-I., Hadlaczky, G., Pettersson, I., Sharp, G. and Ringertz
N. R. (1986) J. Cell. Biol., 102, 137- 144.
93. Carter, K. C., Bowman, D., Carrington, W., Fogarty, K., McNeil, J. A., Fay, F.
S. and Lawrence,J. B. (1993) Science, 259, 1330 - 1334.
94. Xing, Y., Johnson, C. V., Dobner, P. R. and Lawrence, J. B. (1993) Science 259,
1326-1330.
95. Ge, H. and Manley, J. L. (1990) Cell, 62, 25-34.
96. Krainer, A. R., Conway, G. C. and Kozak, D. (1990b) Cell, 62, 35 -42.
97. Fu, X.-D. and Maniatis, T. (1990) Nature, 343, 437-441.
98. Zeitlin, S., Wilson, R. C. and Efstratiadis, A. (1989) J. Cell. Biol., 108, 765777.
99. Roth, M. B., Murphy, C. and Gall, J. G. (199O)J. Cell. Biol., 111, 2217-2223.
100. Roth, M. B., Zahler, A. M. and Stolk, J. A. (1991)J. Cell. Biol., 115, 587-596.
101. Blencowe, B. J., Nickerson, J. A., Issner, R., Penman, S. and Sharp, P. A.
(1993) in preparation.