Download Vertebrates Alternative Adaptive Immunity in Jawless

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Phagocyte wikipedia , lookup

T cell wikipedia , lookup

Drosophila melanogaster wikipedia , lookup

Immune system wikipedia , lookup

Lymphopoiesis wikipedia , lookup

Molecular mimicry wikipedia , lookup

Psychoneuroimmunology wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Cancer immunotherapy wikipedia , lookup

Adaptive immune system wikipedia , lookup

Innate immune system wikipedia , lookup

Adoptive cell transfer wikipedia , lookup

Immunosuppressive drug wikipedia , lookup

X-linked severe combined immunodeficiency wikipedia , lookup

Immunomics wikipedia , lookup

Transcript
Alternative Adaptive Immunity in Jawless
Vertebrates
Brantley R. Herrin and Max D. Cooper
This information is current as
of June 15, 2017.
Subscription
Permissions
Email Alerts
This article cites 37 articles, 17 of which you can access for free at:
http://www.jimmunol.org/content/185/3/1367.full#ref-list-1
Information about subscribing to The Journal of Immunology is online at:
http://jimmunol.org/subscription
Submit copyright permission requests at:
http://www.aai.org/About/Publications/JI/copyright.html
Receive free email-alerts when new articles cite this article. Sign up at:
http://jimmunol.org/alerts
The Journal of Immunology is published twice each month by
The American Association of Immunologists, Inc.,
1451 Rockville Pike, Suite 650, Rockville, MD 20852
Copyright © 2010 by The American Association of
Immunologists, Inc. All rights reserved.
Print ISSN: 0022-1767 Online ISSN: 1550-6606.
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
References
J Immunol 2010; 185:1367-1374; ;
doi: 10.4049/jimmunol.0903128
http://www.jimmunol.org/content/185/3/1367
Alternative Adaptive Immunity in Jawless Vertebrates
Brantley R. Herrin and Max D. Cooper
J
awless vertebrates (agnathans) have been a focal point in
the search for the evolutionary origins of adaptive
immunity because of their unique position in chordate
phylogeny (1). The first agnathans appeared in the Cambrian
period .500 million years ago. However, lamprey and hagfish
are the only groups of agnathans that have survived to the
present. Prior to the agnathans, invertebrate and early chordate
species relied on germline-encoded molecules for detection of
infectious microorganisms and used phagocytes as the primary
immune effector cell type (2). All of the surviving vertebrates
with jaws (gnathostomes) share a complex anticipatory
immune system that features a repertoire of highly diverse Ag
receptors that are generated by somatic rearrangement of Ig V,
D, and J gene segments during the development of clonally
diverse lymphocytes.
In the 1960s and early 1970s, hints of an agnathan adaptive
immune system were provided by experiments in which
immunizationoflampreyswithparticulateAgs,suchasheat-killed
Brucella abortus, bacteriophage f2, or mammalian erythrocytes,
induced the production of specific Ab-like agglutinins (3–6).
Department of Pathology and Laboratory Medicine, Emory Vaccine Center, Emory
University School of Medicine, Atlanta, GA 30322
Received for publication February 16, 2010. Accepted for publication May 10, 2010.
This work was supported by National Institutes Health Grant 5R01AI072435-05 (to
M.D.C.) and the Georgia Research Alliance.
Address correspondence and reprint requests to Dr. Brantley R. Herrin, Department of
Pathology and Laboratory Medicine, Emory Vaccine Center, Emory University School
of Medicine, Room 405, 1462 Clifton Road NE, Atlanta, GA 30322. E-mail address:
[email protected]
www.jimmunol.org/cgi/doi/10.4049/jimmunol.0903128
In addition to these humoral immune responses, delayed-type
reactions to tuberculin and accelerated second set allograft
rejection were suggestive of cell-mediated immunity (3). Despite
the convincing evidence for Ag-specific agglutinins, biochemical
evidence for Ig was not obtained, and the molecular basis for the
lamprey humoral immune response remained elusive (5–8).
Sea lamprey development and anatomy
Sea lamprey (Petromyzon marinus) is the most studied jawless
vertebrate species; larvae hatch from fertilized eggs in inland
streams and burrow into the muddy streambed, where they filter
feed on plankton. The wormlike larvae lack teeth and eyes and
can grow up to 20 cm in length. After 3–14 y, the larvae metamorphose into parasitic adults that resemble eels and migrate
to the sea. Adult sea lampreys develop their eyes and rows of
teeth arranged in concentric circles, which they use to attach to
fish and feed on their body fluids. Adult sea lampreys grow up to
100 cm in length and weigh ∼1 kg. After 12–18 mo at sea, the
lampreys return to freshwater streams to spawn and die shortly
after mating. In the laboratory, lamprey larvae can be housed for
.1 y in aerated, sand-lined aquariums maintained at 15–25˚C
and fed brewer’s yeast. Adult sea lamprey can also be maintained
in aquaculture, but extended culture is more cumbersome
because live fish are required for feeding.
Lampreys lack lymph nodes or a histologically recognizable
thymus, but lymphoid cells are present in their blood and
tissues including the typhlosole (an invagination of the dosal
epithelial wall of the intestine), kidneys, and gills (9, 10). The
typhlosole is the earliest site of hematopoiesis in lamprey
larvae and is sometimes referred to as the spleen (9). During
metamorphosis, the typhlosole regresses, and hematopoiesis
shifts to the dorsal fat body located above the spinal cord. The
intertubular spaces of the anterior kidney are populated by
lymphoid and myeloid cells, and the blood vessels of the gills
contain many lymphocyte-like cells. In addition to nucleated
erythrocytes and thrombocytes, the blood contains cells that
are morphologically similar to the lymphocytes, plasma cells,
monocytes, and granulocytes of jawed vertebrates (9–12).
Discovery of variable lymphocyte receptors
The identification of a genetic basis for adaptive immunity
in jawless vertebrates began with transcriptome analysis of
Abbreviations used in this paper: AID, activation-induced cytidine deaminase; CDA,
cytosine deaminase; CP, connecting peptide; HEL, hen egg lysozyme; HP, hydrophobic
peptide; LRR, leucine-rich repeat; LRR1, the first 24-residue leucine-rich repeat;
LRRCT, C-terminal leucine-rich repeat; LRRNT, N-terminal leucine-rich repeat;
LRRV, variable leucine-rich repeats; LRRVe, end leucine-rich repeat; SP, signal peptide;
VLR, variable lymphocyte receptor.
Copyright Ó 2010 by The American Association of Immunologists, Inc. 0022-1767/10/$16.00
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
Jawless vertebrates use variable lymphocyte receptors
(VLRs) that are generated by RAG-independent combinatorial assembly of leucine-rich repeat cassettes for Ag
recognition, instead of the Ig-based Ag receptors used by
jawed vertebrates. The VLR genes encode for crescentshaped proteins that use variable b-strands and a
C-terminal loop to bind to Ags rather than the six
CDR loops used by BCRs and TCRs. VLR mAbs have
been isolated recently, which enabled the structure of
VLR–Ag complexes to be defined. The jawless vertebrate adaptive immune system has many similarities
to the Ig-based system of jawed vertebrates, including
the compartmentalized development of B-like and
T-like lymphocyte lineages that proliferate and differentiate into VLR-secreting plasmacytes and proinflammatory cytokine-producing cells in response to Ags. The
definition of common features of the VLR-based and
Ig-based systems offers fresh insight into the evolution
of adaptive immunity. The Journal of Immunology,
2010, 185: 1367–1374.
1368
Somatic generation of VLR diversity
After the discovery of a VLR gene in lamprey, two VLR genes
(VLRA and VLRB) were identified in a subsequent analysis of
hagfish (14). An additional VLR gene with homology to hagfish
VLRA was later found in lamprey (15), and the original
lamprey VLR gene was renamed VLRB because of its similarity
to hagfish VLRB. The germline VLRA and VLRB genes are
comprised of two exons, with the first exon encoding only a
portion of the 59 untranslated region (12, 14, 15). The second
exon contains the rest of the 59 untranslated region, a SP, a 59
portion of the LRRNT, a 39 portion of the LRRCT, and the
stalk region. For hagfish VLRA and VLRB and for lamprey
VLRA, the 59 LRRNT sequence is separated from the 39
LRRCT sequence by a short noncoding intervening sequence
that lacks canonical splice donor and acceptor sites (Fig. 1B).
The lamprey VLRB gene is more complex. The 59 LRRNT
coding sequence is separated from a segment encoding a 59
portion of LRRCT by a 5.8-kbp intervening sequence, and a
second 6.4-kbp intervening sequence separates the 59 LRRCT
sequence from the 39 LRRCT and the stalk region (12). The
germline configuration of the VLR genes in all cell types except
lymphocytes lacks direct coding sequences for LRR1, LRRV,
LRRVe, and LRRCP, as well as partial coding sequences for
LRRNT and LRRCT (12, 14, 15). Remarkably, the incomplete germline VLR genes are flanked by an extensive array of
partial LRR gene cassettes (12, 14–17). In lamprey, 393
VLRA-related LRR cassettes and 454 VLRB-related LRR
cassettes have been characterized within a continuous locus of
∼2 mega bp surrounding their respective VLR gene (15).
The VLR gene assembly mechanism has been characterized
by aligning genomic LRR cassettes to mature VLR sequences
and by capturing VLR assembly-intermediate genes that have
LRR cassette inserts, but which have not completed the
assembly process and still retain portions of the intervening
sequence (15–18). These analyses have shown that the lamprey
VLR genes are assembled in a stepwise, piecewise manner in
which the intervening sequence is replaced by random and
sequential copying of the flanking LRR cassettes into the
incomplete VLR gene (Fig. 1C). The assembly process can be
initiated at either the 59 LRRNT or 39 LRRCT ends and
appears to be directed by short stretches of nucleotide
homology (10–30 bp) between donor and acceptor sequences
that guide the copying of flanking LRR segments into the
germline gene (15–17). The VLR gene assembly mechanism
has been described as gene conversion-like (15, 16, 19) because
of the nonreciprocal insertion of LRR cassettes into the intervening sequence of the germline gene, or copy choice (17, 18)
because short stretches of sequence homology are used to target
donor LRR sequences, thereby resembling the mechanism used
by yeast for mating-type switching.
VLR gene assembly occurs on only one allele at a time, a
strategy that is reminiscent of the allelic exclusion of BCR and
TCR gene assembly in jawed vertebrates (16, 17). VLR
assembly is primarily monoallelic, but diallelic assembly has
been detected in a small percentage of hagfish lymphocytes (7%
VLRA, 4% VLRB) (18). In most cases, one of the diallelicly
assembled VLR genes contained base deletions or additions that
resulted in frame-shift mutations, which rendered the gene
nonfunctional. A computational assessment of mature VLRB
gene sequences and the LRR cassettes used to assemble them
predicted a potential repertoire of .1014 distinct Ag receptors,
which is comparable to the theoretical Ab repertoire diversity of
mammalian adaptive immune systems (16).
In contrast to the Ig V, D, and J gene segments in jawed
vertebrates, the LRR cassettes are not flanked by recombination
signal sequences, and RAG1 and 2 have not been identified in
lampreys or hagfish (11–13, 16). Avian and some mammalian
species use a gene conversion mechanism catalyzed by activationinduced cytidine deaminase (AID) to generate Ab diversity (20–
22). Two members of the AID-apolipoprotein B mRNA editing
catalytic component family of DNA cytosine deaminases (CDA)
have been identified in lamprey, and these may be involved in
VLR assembly and diversification (15). CDA1 expression is
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
lymphocyte-like cells isolated from the typhlosole of lamprey
larvae (11, 13). The light-scattering characteristics of these
lymphocyte-like cells were similar to mouse lymphocytes,
which allowed their enrichment by flow cytometry-based cell
sorting (11). The lamprey lymphocyte-like cells were morphologically similar to mammalian lymphocytes by light and
electron microscopy. More than 8000 clones were sequenced
from a cDNA library prepared from the sorted lymphocyte-like
cells, and a basic local alignment search tool analysis of the
resulting expressed sequence tag sequences was conducted to
determine the identity of the orthologous genes (11, 13).
Lamprey genes with homology to transcription factors (Spi B
and Ikaros) and signal transducing molecules (CD45 and B cell
adaptor protein) that are important for lymphocyte development were present in the cDNA library. Nevertheless, the cardinal elements of adaptive immunity, namely Ig, TCR, RAG1
and 2, and MHC class I and II, were conspicuously absent.
In a revised strategy, lampreys were injected with an Ag and
mitogen mixture that induced lymphoblastoid transformation
of lymphocyte-like cells (12). The activated lymphoblastoid
cells in blood samples were sorted by flow cytometry, and a
cDNA library was prepared by subtracting lamprey myeloid
cell and erythrocyte cDNA from lymphocyte-like cDNA. Of
the .1500 subtracted cDNA clones that were sequenced, 239
uniquely diverse leucine-rich repeat (LRR)-containing transcripts were characterized that were exclusively expressed by
lymphocytes, with highest expression being seen in the lymphoblasts. Due to the sequence diversity and lymphocyterestricted expression, these receptors were named variable
lymphocyte receptors (VLRs). Each VLR transcript encoded
an invariant signal peptide (SP) followed by highly variable
LRR modules: a 27–34 residue N-terminal LRR (LRRNT),
the first 24-residue LRR (LRR1), up to eight 24-residue variable LRRs (LRRV), one 24-residue end LRRV (LRRVe), one
16-residue connecting peptide LRR (LRRCP), and a 48–63
residue C-terminal LRR (LRRCT) (Fig. 1A). Southern blotting of lamprey genomic DNA with probes specific for invariant 59 and 39 portions of the VLR sequences revealed that the
highly variable VLRs were encoded by a single VLR gene (12).
The germline VLR gene was incomplete in that it consisted of
sequences encoding for invariant N- and C-terminal portions
of VLRs separated by noncoding intervening sequences.
However, in lymphocyte genomic DNA, the noncoding
intervening sequences were replaced by LRR-encoding
sequences to form a functional, mature VLR gene. All of the
data pointed to the astonishing conclusion that an alternative
form of adaptive immunity comprised of LRR-based Ag
receptors evolved before our Ig-based system.
BRIEF REVIEWS: ALTERNATIVE ADAPTIVE IMMUNITY
The Journal of Immunology
1369
detected in VLRA+ cells, whereas CDA2 expression is restricted
to VLRB-expressing lymphocytes (23). CDA1 was shown to be
highly mutagenic and increased the rate of intragenic mitotic
recombination in yeast (15). These findings favor the conjecture
that CDA1 may catalyze VLRA gene assembly by inducing DNA
strand breaks that allow for homology-based pairing between
donor LRR cassettes and the VLRA gene, whereas CDA2 may
play a similar role in VLRB gene assembly.
VLRB mAbs
The tremendous sequence diversity generated by VLR
assembly suggested they could be useful biomedical reagents if
Ag-specific clones could be isolated. This achievement was
hampered initially because of technical limitations, including
the lack of a cell culture system or a means to immortalize
primary lamprey lymphocytes. However, VLRs are readily
amenable to rDNA techniques, because they are composed of a
single polypeptide, and two methods have been developed to
produce VLR mAbs.
In the first method, VLRB cDNA libraries were prepared
from the lymphocytes of lampreys immunized biweekly for
6 wk with Bacillus anthracis exosporium or human O type
erythrocytes (24). Individual clones from the VLRB cDNA
library were transiently transfected into HEK-293T cells,
which secreted the VLRB proteins as multivalent Abs into the
tissue culture supernatant. The VLRB-containing supernatants
were then screened for Ag-binding by ELISA, immunofluorescence staining, or agglutination assays. VLRB mAbs
were isolated against BclA, the major spore coat protein of
B. anthracis (25), and the H-trisaccharide of human erythrocytes. Fourteen VLRB clones specific for BclA were isolated
out of the 212 tested clones (6.6%) (24). Although all of the
VLRB clones were unique, they shared notable sequence similarity, even in hypervariable positions. All but one of the clones
contained the same number of LRR modules, whereas nonbinding VLRB clones from the same library varied greatly in
the number of LRR modules. VLRB mAbs were highly specific
as evidenced by their ability to discriminate B. anthracis from
Bacillus cereus T spores that differ in BclA sequence at 14 of 134
amino acids, only 9 of which are solvent exposed. VLRB mAbs
isolated using this method were of low affinity (micromolar
KD) but displayed remarkable avidity to Ags with repetitive
epitopes due to their multivalency. For example, an anti-BclA
VLRB mAb (VLR4) agglutinated B. anthracis spores at a 1000fold higher dilution than a corresponding high-affinity mouse
IgG mAb for the same Ag (24).
The VLRB mAbs performed as well as Ig-based Abs did in a
variety of Ab-based assays including ELISA, immunofluorescence staining, and Western blotting. rVLRB mAbs purified
by Ag affinity chromatography required extreme conditions
for elution (pH .11), yet the Abs retained the ability to bind
to Ag despite the harsh elution conditions (24). rVLRB Abs
are also remarkably stable at a variety of temperatures. They
have been stored at 4˚C (either purified or in crude tissueculture supernatant) for .2 y, room temperature for .1 mo, and
56˚C for 36 h without loss of Ag-binding ability. Incubation
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
FIGURE 1. VLR gene assembly. A, Mature VLR genes consist of coding sequences for a SP, an LRRNT, an LRR1, up to eight LRRV cassettes, an LRRVe, a
CP LRR, an LRRCT, an invariant stalk region, and a C-terminal HP, which is preceded by a GPI cleavage site (arrow). B, Two germline VLR genes, VLRA and
VLRB, are present in hagfish and lamprey. The germline VLR genes are incomplete and consist of invariant regions encoding for portions of the LRRNT and
LRRCT separated by noncoding intervening sequences. C, The germline VLR genes are flanked by hundreds of LRR cassettes. During lymphocyte development,
the noncoding intervening sequence is replaced by random and sequential copying of sequences from the flanking LRR cassettes into the germline VLR gene. VLR
gene assembly is initiated at either the LRRNT or LRRCT end and proceeds in a stepwise manner that is guided by short stretches of sequence homology between
the donor and acceptor LRR sequences until a mature VLR gene is generated. Illustrations not drawn to scale. CP, connecting peptide; HP, hydrophobic peptide;
SP, signal peptide.
1370
FIGURE 2. Ag recognition by VLRB Abs.
A, VLR sequence diversity is concentrated on
the concave surface that is composed of
b-strands contributed by LRRNT, LRR1,
LRRV(s), LRRVe, and LRRCP and a variable
loop encoded by LRRCT. B–E, Crystal
structures have been solved for VLRB mAbs in
complex with the H trisaccharide (B) (Brookhaven Protein Data Bank [PDB] code: 3E6J)
and HEL (C) (PDB code: 3G3A). The
LRRCT loop of lamprey VLRB (C, red loop)
and the H3 loop of camel VHH Abs (D, red
loop) (PDB code: 1MEL) are inserted into the
HEL catalytic cleft, whereas VHVL Abs (E)
(PDB code: 1YQV) bind to relatively flat
epitopes away from the catalytic cleft (red
arrow).
was isolated using this approach with 1300-fold higher affinity
(KD = 119 pM from 155 nM) than the parent clone (19). These
data demonstrate that there are no inherent limitations that
prevent VLRs from reaching affinities comparable to mammalian IgG Abs and that a limited number of mutations are
required to obtain high affinity for both types of Ab.
Structural basis for VLR Ag recognition
Crystal structures have been reported for hagfish VLRA and
VLRB without Ag and for lamprey VLRB in complex with Ags
(28–30). VLRA and VLRB encode for crescent-shaped, solenoidal proteins that resemble a cupped hand with parallel
b-strands lining the palm. VLRB contains a thumblike Cterminal variable loop that projects back toward the palm of
the hand. The overall VLR structure is typical of LRR
receptor family members in that the “LxxLxLx” motifs in
LRR1, LRRV, LRRVe, and LRRCP form parallel b-strands
on the concave surface (Fig. 2A). The structure is stabilized by
conserved leucines and phenylalanines that form the hydrophobic core of the solenoid. The hydrophobic core is capped
at the N terminus and C terminus by LRRNT and LRRCT
modules. The LRRNT module contains a b-hairpin structure
in both VLRA and VLRB. The LRRCT is comprised of an
a-helix at the base of the solenoid and a loop positioned on
the concave surface just below the LRRCP b-strand. The
LRRCT loop in VLRB varies greatly in length and sequence
composition and often forms an extended loop that projects
back toward the concave surface (16, 29, 30). Both LRRNT
and LRRCT contain four conserved cysteines that form two
disulfide bridges that further stabilize the structure (28–30).
The VLRA and VLRB sequence diversity is primarily concentrated in the b-strands and the LRRCT loop on the
concave surface that combine to form a continuous hypervariable surface (12, 16, 28–30).
A crystal structure has been reported for a VLRB mAb
(RBC36) that specifically binds to the H-trisaccharide (29).
The H-trisaccharide contacted amino acid residues on the
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
at .70˚C for 1 h was required to denature the VLRB Abs and
prevent Ag binding (B.R. Herrin, unpublished observations).
Molecular weight analysis of monomeric and multimeric
VLRB Abs indicated that the multimers are composed of 8–
10 identical subunits, and the dimeric forms observed under
partial reducing conditions suggested that the multimers are
composed of four to five dimeric subunits (10, 24). Visualization of an rVLRB Ab by electron microscopy revealed that
the pentameric and tetrameric pairs of the VLRB solenoids
connected at a central point by the thin and highly flexible
stalk region (24).
A high-throughput yeast display library screening method has
recently been developed to isolate VLRB mAbs (19). VLRB
proteins were expressed on the surface of yeast by fusion to the
yeast flocculation protein Flo1p, which contains a stalklike
structure and a C-terminal GPI linkage (26, 27). Ag-specific
clones were isolated by sorting the VLRB-expressing yeast that
bound to fluorescently labeled soluble Ags, including hen egg
lysozyme (HEL), R-PE, b-galactosidase, cholera toxin subunit
B, and blood group A and B trisaccharides (19). The yeast
display method allowed for greater breadth of screening, which
enabled VLRB Ag-specific clones to be isolated from both
immunized and nonimmunized libraries and for the selection of
higher affinity clones by in vitro mutagenesis. Three HEL-specific VLRB clones with high nanomolar affinities (KD = 455–
117 nM) were isolated from a VLRB cDNA library constructed
from the lymphocytes of HEL-immunized lamprey larvae.
Although high-affinity VLRB clones could not be isolated from
the HEL-immunized cDNA library, higher affinity mutants
(KD = 55–4.3 nM) of a low-affinity anti-HEL clone, VLRB.2D
(KD = 455 nM), were selected from a mutant library generated
by error-prone PCR (19). Notably, all of the high-affinity
mutants contained substitutions in the hypervariable LRRCT
loop. This finding suggested a strategy of selecting for highaffinity mutants by swapping the LRRCT region of a lowaffinity anti-HEL clone with LRRCT regions PCR amplified
from a large pool of VLRB cDNAs. Remarkably, a mutant clone
BRIEF REVIEWS: ALTERNATIVE ADAPTIVE IMMUNITY
The Journal of Immunology
The VLR stalk region
An invariant threonine/proline-rich stalk region that is germline
encoded connects the C terminus of the LRR solenoid to the cell
membrane (12). Crystal structure data are not available for the
stalk region, but computer modeling predicts that it forms an
extended unstructured region with proline residues functioning to prevent secondary structure formation. The stalk regions
of VLRA and VLRB end with a hydrophobic tail that fits
consensus sequences of GPI-linked receptors (12, 14, 15).
However, only the GPI-linkage of lamprey VLRB has been
verified experimentally (12). Interestingly, VLRA and VLRB do
not encode for cytoplasmic amino acids, a finding that implies
VLR signal transduction is mediated by transmembrane coreceptors that function analogously to Iga/Igb and CD3/
z-chain in jawed vertebrates. In lampreys, VLRB is secreted
into the blood as a disulfide-linked, multivalent Ab, whereas
VLRA is expressed only as a cell-surface receptor (10, 16, 23).
The hydrophobic tail downstream of the lamprey VLRB GPIcleavage site is unusual in that it is rich in cysteine residues that
are not present in VLRA or hagfish VLRB (12, 14, 15, 24).
VLRB cDNAs expressed in mammalian cell lines, such as
HEK-293T cells, are secreted as disulfide-linked multimers
similar to VLRB Abs in lamprey plasma (24). Deletion of the
cysteine-rich hydrophobic tail of the stalk region results in
secretion of monomeric VLRB by transfected HEK-293T cells.
This suggests a mechanism for switching to VLRB secretion by
modulating GPI addition.
Tissue distribution of lamprey lymphocytes
The development of Abs specific for invariant portions of the
VLRA and VLRB receptors allowed the cellular basis of the
lamprey adaptive immune response to be characterized. Surface expression of VLRA and VLRB was detected on distinct
lymphocyte populations by flow cytometry (23). Analysis of
the VLRA and VLRB genomic loci in cells sorted for VLRA
and VLRB surface expression revealed that VLRA gene
assembly only occurs in the VLRA+ cells, and VLRB gene
assembly only occurs in VLRB+ cells. Furthermore, CDA1
expression was restricted to VLRA+ cells, whereas CDA2
expression was detected only in VLRB+ cells (23). The
restricted expression of CDA1 and CDA2 may account for
the mutually exclusive assembly of VLRA and VLRB genes.
The two lineages of lamprey lymphocytes have distinct tissue
distribution patterns, with VLRB cells being the dominant
population in the blood and kidneys, where they outnumber
VLRA cells by an 8:1 ratio. The VLRB to VLRA cell ratio is
2:1 in the typhlosole, and VLRB and VLRA cells are present
in equivalent numbers in the gill region (23). VLRB+ cells are
found by immunohistochemical analysis to be located within
the blood vessels of the gills, in the interstitial spaces between
renal tubules of the anterior kidney, and interspersed among
hematopoietic cells in the typhlosole (10).
Characterization of B-like and T-like lymphocyte populations
Lamprey VLRB lymphocytes have many morphological and
functional similarities to the B cells of jawed vertebrates in
keeping with their function as the humoral arm of the lamprey
adaptive immune system. Immunization of lampreys with
bacteria, viruses, or mammalian cells induces VLRB lymphocytes to undergo lymphoblastoid transformation, proliferation,
and differentiation into plasmacytes that secrete their Ag
receptors as multivalent VLRB Abs (Fig. 3A) (10). In naive
lamprey larvae, ,1% of the VLRB cells in blood are dividing
versus 3–5% of the circulating VLRB cells 14 d postimmunization with B. anthracis exosporium (23). After booster
immunization with B. anthracis exosporium, Ag-specific
VLRB-secreting cells can be detected in the blood, kidneys,
and, to a lesser extent, the typhlosole (10). Electron microscopy
analysis indicates that the VLRB Ab-secreting cells have
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
b-strands of LRR1, LRRV1, LRRV2, LRRV3, LRRVe, LRRCP, and LRRCT but not the LRRNT of RBC36 (Fig. 2B).
The specificity of RBC36 for the H-trisaccharide was primarily mediated by hydrogen bonds contributed by Asp103
of LRRV2 and Asp152 and Gln153 of LRRVe. The H-trisaccharide interaction was also stabilized by van der Waals
contacts contributed by eight other residues located on the
b-strands. In addition to the b-strands, a tryptophan residue
located at the tip of the LRRCT loop contributed a key
interaction by stacking its indole group parallel to the plane of
the galactose ring of the H-trisaccharide (29). The concave
surface area of RBC36 was estimated to be ∼1720 Å2, which
is much larger than the ∼700–1000 Å2 Ag binding site of a
typical mammalian Ab. However, the relatively small H-trisaccharide buried only ∼300 Å2 of the VLRB surface.
A crystal structure has been solved for another VLRB mAb
(VLRB.2D) in complex with a protein Ag, HEL (30). The
interaction is mediated by 19 VLRB.2D residues forming
contacts with 20 residues on the surface of HEL. Similar to
the RBC36-H-trisaccharide complex, HEL makes contacts
with amino acids in LRR1, LRRV, LRRVe, LRRCP, and
LRRCT of VLRB.2D, but not the LRRNT (29, 30). The
most notable feature of the structure is the insertion of the
LRRCT loop into the catalytic cleft of HEL, where it forms a
hydrogen bond with the catalytic residue (Fig. 2C) (30).
Single-chain Abs, camelid VHH (31) and shark IgNAR (32),
recognize almost the same epitope as VLRB.2D by inserting
one of their Ig loops into the catalytic cleft of HEL (Fig. 2D).
In contrast, VHVL mouse Abs that have been crystallized with
HEL invariably bind to relatively flat surfaces away from the
catalytic cleft (Fig. 2E) (33). An in vitro affinity-matured
mutant (VLRB.2DMut13) with ∼10-fold higher affinity for
HEL was also crystallized (30). Superposition of the highaffinity mutant with wild-type VLRB.2B revealed that the
mutations did not cause substantial structural changes. Surprisingly, the two residues mutated in the LRRCT of
VLRB.2DMut13 did not make contact with HEL, and the
total surface area buried by the VLRB.2DMut13-HEL complex was ∼100 Å2 less than the wild-type complex. The
greater affinity of VLRB.2DMut13 was the result of improved
shape and electrostatic complementarity to HEL rather than
formation of additional hydrogen bonds or van der Waals
contacts (30). To examine the role of conformational changes
in Ag binding by VLRB, the structures of bound and
unbound VLRB.2D were compared by superposition. The
concave surface of VLRB.2D maintained almost the same
conformation after HEL binding; however, a hinge movement of ∼1 Å occurred in the LRRCT loop in response to the
HEL interaction (30). The structural rigidity of the LRR
b-strands may make VLRB less prone to induced-fit conformational changes that can lead to nonspecific Ag binding
by the CDR loops of Ig-based Abs.
1371
1372
It is not yet clear whether VLRA cells primarily recognize
native or processed Ags. Yeast display experiments have
identified one example of VLRA clones that can bind to HEL
with high affinity (KD= 26–0.27 nM) (19). Thirteen unique
HEL-specific VLRA clones were isolated from a single animal
that differed from each other at only 15 out of 244 aa positions. The high degree of sequence similarity between the
clones raises the intriguing possibility that the clones were
derived from a single lymphocyte precursor that diversified
its receptor by somatic mutation, perhaps catalyzed by the
AID-like enzyme CDA1 (15, 19). In contrast, VLRA cells
that bind bacteria could not be demonstrated pre- or postimmunization, whereas the numbers of bacteria-binding
VLRB cells increased postimmunization (10, 23). If the
VLRA cells recognize native Ags, like gd T cells, they may
function to coordinate and regulate the immune response via
cytokines and costimulatory molecules. If VLRA cells instead
detect processed Ags displayed on the surface of infected cells,
like ab T cells, they may function to monitor intracellular
compartments that are inaccessible to VLRB Abs. MHC
receptors have not been found in EST sequences (11–13) or
in draft sequences of the lamprey genome (B.R. Herrin,
unpublished observations). Therefore, if VLRA cells detect
processed Ags, the peptides are likely presented by a nonMHC receptor(s) that arose by convergent evolution. Genes
with homology to a TAP-like peptide transporter and the
LMP7/X proteasome subunit have been characterized, which
could suggest the possibility that lamprey cells have the
capacity to generate peptides and transport them for loading
onto a putative receptor for presentation (35, 36).
Differential gene expression by VLRA and VLRB lymphocytes
VLRA and VLRB lymphocytes have been shown to have different gene expression profiles by quantitative analysis of the
transcripts for a selected panel of genes (23). VLRA lymphocytes express a number of genes that are associated with T cell
development in jawed vertebrates. These include several transcription factors (GATA2/3, c-Rel, aryl hydrocarbon receptor,
and BCL11b), the T cell fate-determining receptor Notch1,
the receptor protein tyrosine phosphatase CD45 that is
essential for TCR signaling, and the chemokine receptor
CCR9 that promotes homing of T cell progenitors to the
thymus, although a thymus equivalent has not yet been identified in lamprey (Fig. 3C). The VLRA cells also express the
FIGURE 3. B-like and T-like lymphocytes in lamprey. A and B, Particulate Ags stimulate lymphoblastoid transformation and proliferation of VLRB+ and
VLRA+ cells. VLRB+ cells differentiate into VLRB Ab-secreting cells that resemble plasma cells in jawed vertebrates (A). In contrast, activated VLRA+ cells
differentiate into proinflammatory cytokine-producing effector cells that do not secrete their Ag receptors. C, VLRA and VLRB cells express many genes that are
required for B cell and T cell development in jawed vertebrates. The reciprocal expression of cytokines and cytokine receptors by VLRB+ and VLRA+ cells suggests
the possibility of functional interactions between the two lymphocyte lineages.
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
extensive cytoplasm and rough endoplasmic reticulum, much
like plasma cells in jawed vertebrate animals.
The VLRB Ab response evolves more slowly than the Ab
responses of mammals and is more similar to the IgM
responses of cartilaginous and bony fishes (10, 34). The peak
Ab levels observed in the lamprey VLRB response are also
lower than in the mammalian Ab response. The highest Agspecific plasma VLRB titer obtained after repeated immunization is ∼1:10,000 in lamprey larvae, whereas mammalian
IgG titers may exceed 1:100,000. Lampreys immunized with
B. anthracis exosporium produced VLRB Abs that reacted
with the BclA spore coat protein, which is also the dominant
Ag recognized by Abs in mice immunized with B. anthracis
exosporium (10, 25). VLRB Abs specific for BclA were
detectable by day 7 postimmunization, and their levels peaked
by day 26, although appreciable anti-BclA titers persisted
beyond day 50 (10). A second immunization with B. anthracis
exosporium induced several-fold higher titers of VLRB Abs
for BclA relative to lampreys given a single immunization.
Similar kinetics were observed for the VLRB Ab response
in lampreys immunized with human O type erythrocytes.
VLRB agglutination titers peaked at day 19 after a single
immunization, and a second immunization resulted in a 20fold enhancement in agglutination titer (10).
VLRA cells also undergo lymphoblastoid transformation
and proliferation postimmunization with viruses, bacteria, and
mammalian cells, but they do not produce soluble VLRA
proteins either pre- or postimmunization with bacteria or
stimulation with the classical T cell mitogen, PHA (Fig. 3B)
(23). Moreover, the VLRA molecules are not secreted by
HEK-293T cells transfected with VLRA cDNAs, whereas
VLRB Abs are secreted by transfected heterologous cells (23,
24). PHA also does not induce VLRA secretion, but it preferentially stimulates vigorous proliferation of VLRA cells,
leading them to become the dominant lymphocyte population by 9 d postinjection into lamprey larvae (23). The
activated VLRA lymphoblasts are more than twice the size of
naive VLRA cells and have expanded cytoplasm, mitochondria, and endoplasmic reticulum. Importantly, the activated
VLRA cells express higher transcript levels for proinflammatory
cytokines, macrophage migration inhibitory factor, and IL-17
(23). These findings suggest that Ag-activated VLRA cells
differentiate into cytokine-secreting effector cells that may
direct the immune response.
BRIEF REVIEWS: ALTERNATIVE ADAPTIVE IMMUNITY
The Journal of Immunology
proinflammatory cytokines, macrophage migration inhibitory
factor, and IL-17, as well as the IL-8R (CXCR2). In contrast,
the VLRB lymphocytes express many genes that are required
for B cell activation and differentiation in jawed vertebrates,
including the protein tyrosine kinase Syk and the B cell adaptor
protein that transduce signals initiated by BCR ligation, the
TNFR superfamily member 14/herpesvirus entry mediator
that binds to LIGHT or B and T lymphocyte attenuator on the
surface of T cells, and the chemokine receptor CXCR4 (23).
Like mammalian B cells, the VLRB lymphocytes also preferentially express several TLRs (TLR2abc, 7, 10) that may synergize with Ag receptor signals to promote activation. One of
the most notable findings for the VLRB cells is their expression
of the chemotactic and inflammatory cytokine IL-8 and the
receptor for IL-17, which complements the expression of IL8R and IL-17 by VLRA cells to suggest the possibility of
cooperative functional interactions between the two lymphocyte lineages.
There is much left to be learned about the adaptive immune
system in agnathans before we can understand the survival
advantage that favored its acquisition, presumably to augment
innate mechanisms of protective immunity against pathogens.
Although the Ig-based and VLR-based adaptive immune
systems in jawed and jawless vertebrates use different genes and
assembly mechanisms, both systems generate diverse repertoires of anticipatory receptors capable of recognizing almost
any Ag through the combinatorial assembly of large arrays of
partial gene segments. The development of clonally diverse
lymphocytes allows for Ag-specific responses and memory,
which are lacking in innate immunity. However, these
advantages must offset the inherent danger associated with the
stochastic assembly of anticipatory receptor genes by effector
lymphocytes, namely the potential to generate receptors with
specificities for self Ags. Although we know that the Ig-based
system has evolved multiple mechanisms to ensure tolerance to
self, the tolerance phenomenon has not yet been explored in
the VLR-based system. It will be fascinating to learn how this
fundamental immunological dilemma is resolved in jawless
vertebrates.
Hybridoma technology enabled the diversity of the mammalian B cell repertoire to be harnessed to produce mAbs with
high affinity and specificity for widespread use as experimental,
diagnostic, and therapeutic reagents. rDNA techniques and Ab
display technologies have facilitated the production of VLRB
mAbs and revealed several advantages, including broad pH and
temperature stability, lack of tolerance to mammalian Ags, and
modular, single peptide composition that make them an
attractive alternative to mammalian VHVL mAbs. Although
artificial alternatives to Ig Abs have been constructed using
protein A domains, ankyrin repeats, fibronectin domains, and
lipocalin scaffolds among others (37), VLRs are the only
naturally occurring Ig alternative for which the Ag recognition
proficiency has been honed by hundreds of millions of years
of natural selection. Only 6 y after their discovery, VLRs
already have demonstrable biotechnology potential, and the
quest to exploit these ancient Abs has just begun.
Disclosures
The authors have no financial conflicts of interest.
References
1. Pancer, Z., and M. D. Cooper. 2006. The evolution of adaptive immunity. Annu.
Rev. Immunol. 24: 497–518.
2. Litman, G. W., L. J. Dishaw, J. P. Cannon, R. N. Haire, and J. P. Rast. 2007. Alternative mechanisms of immune receptor diversity. Curr. Opin. Immunol. 19: 526–534.
3. Finstad, J., and R. A. Good. 1964. The Evolution of the Immune Response. III.
Immunologic Responses in the Lamprey. J. Exp. Med. 120: 1151–1168.
4. Boffa, G. A., J. M. Fine, A. Drilhon, and P. Amouch. 1967. Immunoglobulins and
transferrin in marine lamprey sera. Nature 214: 700–702.
5. Marchalonis, J. J., and G. M. Edelman. 1968. Phylogenetic origins of antibody
structure. 3. Antibodies in the primary immune response of the sea lamprey, Petromyzon marinus. J. Exp. Med. 127: 891–914.
6. Pollara, B., G. W. Litman, J. Finstad, J. Howell, and R. A. Good. 1970. The
evolution of the immune response. VII. Antibody to human “O” cells and properties of the immunoglobulin in lamprey. J. Immunol. 105: 738–745.
7. Litman, G. W., F. J. Finstad, J. Howell, B. W. Pollara, and R. A. God. 1970. The
evolution of the immune response. 3. Structural studies of the lamprey immuoglobulin. J. Immunol. 105: 1278–1285.
8. Ishiguro, H., K. Kobayashi, M. Suzuki, K. Titani, S. Tomonaga, and Y. Kurosawa.
1992. Isolation of a hagfish gene that encodes a complement component. EMBO J.
11: 829–837.
9. Amemiya, C. T., N. R. Saha, and A. Zapata. 2007. Evolution and development of
immunological structures in the lamprey. Curr. Opin. Immunol. 19: 535–541.
10. Alder, M. N., B. R. Herrin, A. Sadlonova, C. R. Stockard, W. E. Grizzle,
L. A. Gartland, G. L. Gartland, J. A. Boydston, C. L. Turnbough Jr., and
M. D. Cooper. 2008. Antibody responses of variable lymphocyte receptors in the
lamprey. Nat. Immunol. 9: 319–327.
11. Mayer, W. E., T. Uinuk-Ool, H. Tichy, L. A. Gartland, J. Klein, and
M. D. Cooper. 2002. Isolation and characterization of lymphocyte-like cells from a
lamprey. Proc. Natl. Acad. Sci. USA 99: 14350–14355.
12. Pancer, Z., C. T. Amemiya, G. R. Ehrhardt, J. Ceitlin, G. L. Gartland, and
M. D. Cooper. 2004. Somatic diversification of variable lymphocyte receptors in the
agnathan sea lamprey. Nature 430: 174–180.
13. Uinuk-Ool, T., W. E. Mayer, A. Sato, R. Dongak, M. D. Cooper, and J. Klein.
2002. Lamprey lymphocyte-like cells express homologs of genes involved in
immunologically relevant activities of mammalian lymphocytes. Proc. Natl. Acad.
Sci. USA 99: 14356–14361.
14. Pancer, Z., N. R. Saha, J. Kasamatsu, T. Suzuki, C. T. Amemiya, M. Kasahara, and
M. D. Cooper. 2005. Variable lymphocyte receptors in hagfish. Proc. Natl. Acad.
Sci. USA 102: 9224–9229.
15. Rogozin, I. B., L. M. Iyer, L. Liang, G. V. Glazko, V. G. Liston, Y. I. Pavlov,
L. Aravind, and Z. Pancer. 2007. Evolution and diversification of lamprey antigen
receptors: evidence for involvement of an AID-APOBEC family cytosine deaminase.
Nat. Immunol. 8: 647–656.
16. Alder, M. N., I. B. Rogozin, L. M. Iyer, G. V. Glazko, M. D. Cooper, and
Z. Pancer. 2005. Diversity and function of adaptive immune receptors in a jawless
vertebrate. Science 310: 1970–1973.
17. Nagawa, F., N. Kishishita, K. Shimizu, S. Hirose, M. Miyoshi, J. Nezu,
T. Nishimura, H. Nishizumi, Y. Takahashi, S. Hashimoto, et al. 2007. Antigenreceptor genes of the agnathan lamprey are assembled by a process involving copy
choice. Nat. Immunol. 8: 206–213.
18. Kishishita, N., T. Matsuno, Y. Takahashi, H. Takaba, H. Nishizumi, and
F. Nagawa. 2010. Regulation of antigen-receptor gene assembly in hagfish. EMBO
Rep. 11: 126–132.
19. Tasumi, S., C. A. Velikovsky, G. Xu, S. A. Gai, K. D. Wittrup, M. F. Flajnik,
R. A. Mariuzza, and Z. Pancer. 2009. High-affinity lamprey VLRA and VLRB
monoclonal antibodies. Proc. Natl. Acad. Sci. USA 106: 12891–12896.
20. Arakawa, H., J. Hauschild, and J. M. Buerstedde. 2002. Requirement of the
activation-induced deaminase (AID) gene for immunoglobulin gene conversion.
Science 295: 1301–1306.
21. Honjo, T., M. Muramatsu, and S. Fagarasan. 2004. AID: how does it aid antibody
diversity? Immunity 20: 659–668.
22. Di Noia, J. M., and M. S. Neuberger. 2007. Molecular mechanisms of antibody
somatic hypermutation. Annu. Rev. Biochem. 76: 1–22.
23. Guo, P., M. Hirano, B. R. Herrin, J. Li, C. Yu, A. Sadlonova, and M. D. Cooper.
2009. Dual nature of the adaptive immune system in lampreys. Nature 459: 796–801.
24. Herrin, B. R., M. N. Alder, K. H. Roux, C. Sina, G. R. Ehrhardt, J. A. Boydston,
C. L. Turnbough Jr., and M. D. Cooper. 2008. Structure and specificity of lamprey
monoclonal antibodies. Proc. Natl. Acad. Sci. USA 105: 2040–2045.
25. Steichen, C., P. Chen, J. F. Kearney, and C. L. Turnbough Jr. 2003. Identification
of the immunodominant protein and other proteins of the Bacillus anthracis exosporium. J. Bacteriol. 185: 1903–1910.
26. Gai, S. A., and K. D. Wittrup. 2007. Yeast surface display for protein engineering
and characterization. Curr. Opin. Struct. Biol. 17: 467–473.
27. Sato, N., T. Matsumoto, M. Ueda, A. Tanaka, H. Fukuda, and A. Kondo. 2002.
Long anchor using Flo1 protein enhances reactivity of cell surface-displayed glucoamylase to polymer substrates. Appl. Microbiol. Biotechnol. 60: 469–474.
28. Kim, H. M., S. C. Oh, K. J. Lim, J. Kasamatsu, J. Y. Heo, B. S. Park, H. Lee,
O. J. Yoo, M. Kasahara, and J. O. Lee. 2007. Structural diversity of the hagfish
variable lymphocyte receptors. J. Biol. Chem. 282: 6726–6732.
29. Han, B. W., B. R. Herrin, M. D. Cooper, and I. A. Wilson. 2008. Antigen recognition by variable lymphocyte receptors. Science 321: 1834–1837.
30. Velikovsky, C. A., L. Deng, S. Tasumi, L. M. Iyer, M. C. Kerzic, L. Aravind,
Z. Pancer, and R. A. Mariuzza. 2009. Structure of a lamprey variable lymphocyte
receptor in complex with a protein antigen. Nat. Struct. Mol. Biol. 16: 725–730.
31. De Genst, E., K. Silence, K. Decanniere, K. Conrath, R. Loris, J. Kinne,
S. Muyldermans, and L. Wyns. 2006. Molecular basis for the preferential cleft
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017
Conclusions
1373
1374
recognition by dromedary heavy-chain antibodies. Proc. Natl. Acad. Sci. USA 103:
4586–4591.
32. Stanfield, R. L., H. Dooley, M. F. Flajnik, and I. A. Wilson. 2004. Crystal structure
of a shark single-domain antibody V region in complex with lysozyme. Science 305:
1770–1773.
33. Bentley, G. A. 1996. The crystal structures of complexes formed between lysozyme
and antibody fragments. EXS 75: 301–319.
34. Dooley, H., and M. F. Flajnik. 2005. Shark immunity bites back: affinity maturation and memory response in the nurse shark, Ginglymostoma cirratum. Eur. J.
Immunol. 35: 936–945.
BRIEF REVIEWS: ALTERNATIVE ADAPTIVE IMMUNITY
35. Kandil, E., C. Namikawa, M. Nonaka, A. S. Greenberg, M. F. Flajnik, T. Ishibashi,
and M. Kasahara. 1996. Isolation of low molecular mass polypeptide complementary DNA clones from primitive vertebrates. Implications for the origin of
MHC class I-restricted antigen presentation. J. Immunol. 156: 4245–4253.
36. Uinuk-ool, T. S., W. E. Mayer, A. Sato, N. Takezaki, L. Benyon, M. D. Cooper,
and J. Klein. 2003. Identification and characterization of a TAP-family gene in the
lamprey. Immunogenetics 55: 38–48.
37. Binz, H. K., P. Amstutz, and A. Plückthun. 2005. Engineering novel binding proteins
from nonimmunoglobulin domains. Nat. Biotechnol. 23: 1257–1268.
Downloaded from http://www.jimmunol.org/ by guest on June 15, 2017