Sei sulla pagina 1di 11

Biochimica et Biophysica Acta 1860 (2016) 333–343

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

journal homepage: www.elsevier.com/locate/bbagen

Gamma crystallins of the human eye lens☆


Venkata Pulla Rao Vendra a, Ismail Khan b, Sushil Chandani c,
Anbukkarasi Muniyandi d, Dorairajan Balasubramanian b,⁎
a
Ophthalmic Molecular Genetics Section, National Eye Institute, Building 5635FL, Room 1S24, 5625 Fishers Lane, Rockville, MD 20852, United States
b
Prof. Brien Holden Eye Research Centre, Hyderabad Eye Research Foundation, L. V. Prasad Eye Institute, Hyderabad 500034 Telangana, India
c
Plot 32, LIC Colony, W Marredpally, Secunderabad 500026, Telangana, India
d
Department of Animal Science, Bharathidasan University, Tiruchirappalli 620 024, Tamil Nadu, India

a r t i c l e i n f o a b s t r a c t

Article history: Background: Protein crystallins co me in three types (α, β and γ) and are found predominantly in the eye, and
Received 17 April 2015 particularly in the lens, where they are packed into a compact, plastic, elastic, and transparent globule of proper
Received in revised form 8 June 2015 refractive power range that aids in focusing incoming light on to the retina. Of these, the γ-crystallins are found
Accepted 19 June 2015 largely in the nuclear region of the lens at very high concentrations (N 400 mg/ml). The connection between their
Available online 25 June 2015
structure and inter-molecular interactions and lens transparency is an issue of particular interest.
Scope of review: We review the origin and phylogeny of the gamma crystallins, their special structure involving
Keywords:
Human gamma crystallins
the use of Greek key supersecondary structural motif, and how they aid in offering the appropriate refractive
Greek key motif index gradient, intermolecular short range attractive interactions (aiding in packing them into a transparent
Structure–function correlation ball), the role that several of the constituent amino acid residues play in this process, the thermodynamic and
Cataractogenic mutations kinetic stability and how even single point mutations can upset this delicate balance and lead to intermolecular
Congenital cataract aggregation, forming light-scattering particles which compromise transparency. We cite several examples of this,
and illustrate this by cloning, expressing, isolating and comparing the properties of the mutant protein S39C of
human γS-crystallin (associated with congenital cataract-microcornea), with those of the wild type molecule.
In addition, we note that human γ-crystallins are also present in other parts of the eye (e.g., retina), where
their functions are yet to be understood.
Major conclusions: There are several ‘crucial’ residues in and around the Greek key motifs which are essential to
maintain the compact architecture of the crystallin molecules. We find that a mutation that replaces even one of
these residues can lead to reduction in solubility, formation of light-scattering particles and loss of transparency
in the molecular assembly.
General significance: Such a molecular understanding of the process helps us construct the continuum of
genotype–molecular structural phenotype–clinical (pathological) phenotype. This article is part of a Special
Issue entitled Crystallin Biochemistry in Health and Disease.
© 2016 Elsevier B.V. All rights reserved.

1. Gamma crystallins: phylogeny, presence in the lens, functions mammalian eye lenses and N 1000 mg/ml in some fish lenses). They
are packed compactly in short range spatial order as in dense liquids
Lens crystallins constitute about 90% of the total soluble protein con- or glasses such that they aid transparency (no scattering) and the ap-
tent of the lens, and 35% of the total lens mass. They are composed of propriate refractive index gradient, high in the nuclear region and
two broad classes: α and βγ. The distribution of the three crystallins lower in the cortical region of the lens. These two features are achieved
in the lens is asymmetric and biphasic [1]. The central portion of the by the unique and stable structure and packing that they have adopted,
lens is rich in β- and γ-crystallins, and the embryonic nuclear region, a classic example of structure dictating function.
from where the lens grows, is particularly rich in γ-crystallins. These Of the three crystallins found in the vertebrate lens (α, β and γ),
proteins are the earliest to be expressed as the lens is formed and gamma crystallins are the smallest and simplest members. Each of
grows, and present in very high concentrations (N400 mg/ml in them has about 175 amino acids in its sequence, with a molecular
weight of about 21 kDa. Each of them is folded in a compact globular
☆ This article is part of a Special Issue entitled Crystallin Biochemistry in Health manner and exists as a monomer. At least six different types of
and Disease. γ-crystallins have been reported so far, namely γA–γF. In addition,
⁎ Corresponding author.
E-mail addresses: pullaraovv@gmail.com (V.P.R. Vendra), ismailkhan.hcu@gmail.com
another one, which used to be called βS-crystallin, is now reclassified as
(I. Khan), sushilchandani@gmail.com (S. Chandani), anbu.blessings@gmail.com γS-crystallin, based on its similarity with the other γ-crystallins. The
(A. Muniyandi), dbala@lvpei.org (D. Balasubramanian). human eye lens is known to contain γC, γD and γS crystallins (γE and

http://dx.doi.org/10.1016/j.bbagen.2015.06.007
0304-4165/© 2016 Elsevier B.V. All rights reserved.
334 V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343

Table 1
Primary structures of human γC-, γD- and γS-crystallins.

An alignment of the amino acid sequences of human γC, γD and γS crystallins. Positions with invariant amino acids are shown in
bold. Serine 39 in γS crystallin is underscored. Colored stretches correspond to the four Greek key motifs.

γF- are pseudogenes in humans). While γC and γD are synthesized any, there. The putative role of lens crystallins, particularly γ-
during the embryonic stage, γS is synthesized post-natally and with de- crystallins, is an area that is drawing considerable current attention.
velopment their relative proportion decreases with concomitant
increase of α- and β-crystallins. 2. Section I: the Greek key motif: role in structure and function
Gamma crystallin is part of the βγ-family and known to have evolved
from primordial sources, namely archaeabacteria, and a description of The primary structural sequences of human γ-crystallins are shown
such microbial βγ-crystallins has been provided recently by Mishra in Table 1.
et al. [2]. The solution state structure of such a primordial βγ-crystallin A common protein structural folding pattern that has been pre-
from Archaea has been published by Barnwal et al. [3]. The native state served through the phylogeny is the so called Greek key motif, a
stability and related properties of the βγ-crystallins have just been super-secondary structural fold that was originally defined and named
reviewed by Serebryany and King (4). How these proteins have evolved by Richardson [8]. The Greek key motif involves a run of four β-sheet
over time in order to play a role in the vertebrate eye lens, and what structures folded in an antiparallel manner to generate a neatly folded
their functions are in and out of the lens have been comprehensively sum- compact domain, as shown in Fig. 1. Several archaeal proteins such as
marized recently by Slingsby, Wistow and Clark [5], and Wistow and Nitrollin or M-crystallin are folded using one such Greek key domain
Slingsby [6]. Given that these comprehensive reviews [2,4–6] have just while the eye lens βγ-crystallins are two-domain structures (with two
appeared, and that the present article is the lone one on γ-crystallins in Greek key motifs per domain), and the protein referred to as mammali-
this special issue, we first present here a ‘review of reviews’ on human an Absent in Melanoma 1 (AIM-1) uses as many as six such domains [2].
γ-crystallins, and follow it up with how mutations in them are associated Indeed, it is this folding using the Greek key motif that renders
with congenital cataracts in humans, an area of special interest to eye care the γ-crystallins with a variety of favorable features and properties
centers such as ours; our center alone treats as many as 1200 such pa-
tients each year. Congenital cataracts are largely genetic in origin (muta-
tions in over 45 genetic loci and over 38 genes in humans [7]), unlike
other forms of cataract where metabolic, environmental and post-
translational modifications can alter the structural and functional proper-
ties of the constituents of the lens; also a large fraction of them are Men-
delian in origin (one gene–one phenotype) so that a genotype–
phenotype correlation is easier. Further, since the detailed molecular
structures of several lens crystallins, particularly the βγ-crystallins are in-
creasingly being unraveled, we can attempt a genotype–molecular struc-
tural phenotype–clinical phenotype correlation.
We are interested here in congenital cataracts associated with
mutations in human γ-crystallins (as many as 30 are reported to date
in γC, γD and γS-crystallins). Since the detailed molecular structures
of these molecules have been solved, making such a genotype–structural
phenotype–clinical phenotype correlation appears to be a possibility.
We have divided this paper into three broad sections. In the first
section, we review the results of a significant number of papers in
the literature which have described the unique structural features of
γ-crystallins, namely, the Greek key motif, and summarize how folding
of the molecule using this motif offers these proteins the right set of
properties needed in the human lens, such as compactness, right
refractive properties, stability and intermolecular packing to generate
a transparent assembly.
In the second section, we show how single point mutations in their se-
quence can lead to perturbations in the structure leading to intermolecu-
lar aggregation and the consequent compromise in lens transparency. As
an example, we present our own ongoing research on the mutant S39C of Fig. 1. The Greek key motif. A: an artistic rendering of the motif; B: a schematic of the ar-
rangement of beta strands in one Greek key domain of the γ-crystallins; C: ribbon diagram
human γS-crystallin, associated with cataract and microcornea in a child. of γS crystallin with the motifs in blue, green, yellow and red, and D: Corey–Pauling–
In the third section, we briefly discuss the presence of crystallins in Koltun space-filling model rendering the same perspective as in Fig. 1C and E rotated
tissues outside the lens and some ongoing research on their role, if 90° for a view from the top.
V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343 335

suited for their role in the lens. The intramolecular packing involving and C-terminal domains fold upon each other, thus making the mole-
the coming together of the N-terminal and C-terminal domains cule a compact globule. Human γD-crystallin is folded in exactly the
generates a compact monomeric globule of about 5 nm in size, same way as human γC-crystallin. In human γS-crystallin, shown in
i.e., Stokes radius of 2.13 nm [9], and a low frictional ratio of about Fig. 1, the sequence 6–44 folds into one Greek key motif, the residues
1.21 (only slightly above the value of 1.12 for a perfect compact smooth 45–87 fold as the second, while the sequence 94–134 forms the third
protein sphere), suggesting that they have low hydration and a low and the stretch 135–177 adopts the last motif. Residues 88–93 form
propensity for sticky interactions with solvent and perhaps, with the connecting peptide that allows the intramolecular coming together
other proteins [10]. These small globules, packed at such high concen- of the N-terminal and C-terminal domains to produce a compact
trations in the lens (400–1000 mg/ml), are seen to exhibit short range globular molecule. (In contrast, the linker region plus N-terminal and
order which is necessary for transparency, particularly in a crowded C-terminal extensions in the β-crystallins together lead to inter-
macromolecular environment as in the lens [11]. Interestingly, it has molecular, rather than intra-molecular, interactions generating dimers
been reported that while interactions between γ-crystallins are weakly and multimers [13,14] rather than monomers like in γ-crystallins.)
attractive, those between β-crystallins appear to be mildly repulsive [12]. Also, while the βγ-crystallins fold using the Greek key motifs,
Table 1 lists the sequences of the three γ-crystallins found in α-crystallin, which belongs to the small heat shock protein family,
humans. Human γC-crystallin, a protein of 174 residues, is folded in does not. It is a multimeric protein with molecular weights as high
four Greek key motifs, residues 2–40 forming the first motif, residues as 800 kDa. The structures and properties of some members of the β-
41–83 forming the second (both in the N-terminal domain of the crystallin family, and those of α-crystallins are dealt with by other
molecule) and residues 88–128 making up the third and 129–171 contributors to this special issue.
making up the fourth and last Greek key fold. Residues 84–87 make The crystal structure of calf lens gamma-II crystallin was first pub-
up the ‘linker region’ which, being a short one, brings the N-terminal lished by Blundell et al. [15] and Wistow et al. [16], and that of human

Fig. 2. Tyrosine corners and domain interface. Fig. 2A: The tyrosine corner domain interface in human γD-crystallin. In the N-terminal domain (left), the side chain hydroxyl group of Y63
forms a H-bond with the backbone oxygen of R59 (residue N-4). In the C-terminal domain, it is Y151 and M147. The inset shows only the N-terminal domain tyrosine corner; the three
intervening residues are in gray. Fig. 2B: Residues involved in stabilizing the domain interface in human γD-crystallin. Three hydrophobic residues from each domain are depicted in the
CPK structural format; the two polar residues of each domain are shown in the ball-and-stick format. The hydrogens have been stripped for clarity, and the carbon atoms in residues in the
N-terminal domain are colored green, while those in the C-terminal domain are in pink.
336 V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343

γD-crystallin by Basak et al. [17]. (Also, the crystal structure of the C- amino acid refractivity values of each constituent residue [23], and not
terminal domain of human γS-crystallin has been determined [18].) necessarily its actual sequence. It is noted that the dn/dc values of
The molecule is seen to pack in a highly symmetric manner, with each double Greek key domains are consistently higher (0.2009 ml/g for
of the two domains as tightly packed β-sheet sandwiches made of the Greek key motifs of crystallins, cf. 0.1899 for other human proteins
Greek key motifs. As Chen et al. [12] point out, this template structure [19]). Likewise, Zhao et al. [20] had earlier argued that “one could spec-
is highly conserved, thanks to several key residues in the sequence ulate that this flexibility of the Greek key motif to accommodate differ-
which play essential roles in the folding pattern, yet tolerates some var- ent residues lends itself well as a stable scaffold for lens proteins,
iations in the non-essential residues in the sequence. The relation be- providing the opportunity of replacing residues with low dn/dc with
tween the role of the Greek key structure and the values and gradients high ones”. It should also be noted that γ-crystallins have the highest
in the refractive index of lens proteins has been drawn attention to in dn/dc values, namely 0.203 ml/g while β-crystallins have the lowest
some recent papers [19,20]. (0.187 ml/g, and α-crystallins in between [24]). This is in keeping
The function of the lens is to focus the light coming through the cor- with the fact that γ-crystallins are the most abundant proteins in the
nea into the retina. This would need not only the appropriate refractive core of the lens where the refractive index is the highest.
index values, but also a refractive index gradient which minimizes Table 1 above lists the amino acid sequences of the three human γ-
spherical aberration. The lens should also be plastic (changeable in crystallins, namely γC, γD and γS. Note the remarkable sequence ho-
shape) and elastic (return to original shape after pressure is released), mology between the three, suggesting as well that they will all be folded
so as to accommodate its size and shape in a manner such that near ob- in much the same manner, which has proved to be the case.
jects as well as far objects should be focused on the retina. While the av- The crystal structural analysis of γ-crystallins shows all nonpolar
erage refractive index of the nuclear region of the porcine lens is about residues in the molecule to be packed in the interior and polar ones
1.45, it decreases to about 1.36 as we move from the central to the cor- on the surface. Solution state studies also bear this out. This also
tical region on the equatorial plane [21]. The macromolecular concen- makes the molecule both thermodynamically and kinetically stable.
tration gradient (very high levels to gradually lower ones) that is seen The intra- and inter-molecular packing features of the molecule (partic-
as we go from the nuclear to the cortical region of the live lens is in ularly human γC, γD and γS), and the roles played by various constitu-
what has been called opto-biological synchrony [22]. The refractive ent residues have been dissected and reported in a series of about 30
index of a protein can be estimated based on the refractive index incre- papers by Jonathan King's group [25] and in about 12 papers by the
ment (dn/dc) values from the amino acid composition, using known Pande group [26] using optical spectroscopic and thermodynamic

Table 2
Mutations reported in human γD-, γC- and γS-crystallins (taken with permission from Vendra et al. [62]).

S. no. Mutation Type of cataract seen Reference Comments

A: Mutations reported in human γD-crystallin


1 R14C Nuclear and perinuclear [45,46] Only surface change; extensive disulfide intermolecular bridges (see ref. 47)
2 R14S Coralliform [48] Slight change in hydrophobic content, opening up a phosphorylation site (no conformational work)
3 P24T Coralliform, cerulean, fasciculiform [49–55] Extensive conformational analysis done by many shows no change in sec/tert structure, minor
surface changes, drop in solubility (see refs. 56–60)
4 A36P Nuclear [61] Greek key compactness distorted, higher hydrophobic exposure upon mutation; solubility drops
(see ref. 62)
5 R37P Nuclear opacity [63] No data yet
6 R37S Birefringent crystals [64] Crystal structure reveals no changes (see ref. 65)
7 W43R Dominant congenital cataract [44] Minor tert struct. change; pI must be different, solubility not affected (see ref. 44)
8 M44V Blue dot opacity [66] No structural data yet
9 Y56X Nuclear [67] Large scale truncation of chain; three Greek key motifs lost
10 R58H Aculeiform [68] H bonding lost, but no other change. pI altered (see ref. 65)
11 G61C Coralliform [69,70] Mutation mainly affected the transition from native to intermediate state but not from intermediate
to unfolded form.
12 R77S Polar coronary cataract [71,72] No change seen in structure upon mutation. pI change? Solubility high (see ref. 62)
13 E107A Nuclear [73,74] Same as above. Solubility high. pI change is seen to lead to heteroaggregation with α-crystallin
(see ref. 62)
14 Y134X Not reported [75] No data, but Greek key 4 lost (see ref. 62)
15 R140X Nuclear [76] Greek key 4 lost; solubility lost; surface exposure extensive (see ref. 62)
16 W157X Nuclear [77] Greek key 4 lost; solubility lost; surface exposure extensive (see refs. 62,78)
17 G165fs Nuclear [79] High hydrophobic exposure; solubility lost (see ref. 62)

B: Mutations reported in human γC-crystallin


1 T5P Coppock-like cataract [68] No data
2 G62fs Zonular pulverulant [80] No data
3 C109X Nuclear cataract [81] Truncation and loss of Greek key
4 S119S Nuclear cataract [67] Polymorphism
5 W157X Nuclear cataract [82] Truncation mutant; data shows solubility loss, exposure of residues to surface (see refs. 62,78)
6 R168W Nuclear cataract [77,83] pI change, as in E107A? (see ref. 62)
7 R48H Nuclear cataract [84] Similar to the γD- mutant
8 G129C Nuclear cataract [85] Tertiary structure appears impaired

C: Mutations reported in human γS-crystallin


1 G18V Progressive cortical cataract [86] Extensive data shows no change in 2° or 3°, but compaction of GK 1 likely to alter (see refs. 87,88)
2 D26G Coppock cataract [61] Very little change in 2° or 3°, solubility high (see ref. 89)
3 S39C Microcornea-cataract [76] Surface hydrophobicity increases, causing the tendency to self aggregate (current study)
4 V42M Autosomal dominant cataract [90] Compact packing of Greek key is distorted (see refs. 91,92)
5 G57W Autosomal dominant pulverulent cataract [93] G highly conserved. Substitution by W partially disrupts the second Greek key motif.

Residue numbers in column 2 are as reported with cited references, regardless of whether methionine1 was counted or not. Note-2: Coralliform: round or elongated processes radiating
out of the center of the lens; cerulean: small bluish dots; fasciculiform: fibrous radiative strands emanating from the center; aculeiform: frosted lens; polar coronary: ring around a clear
central lens, slowly progressive cortical opacity; Coppock-like: bilateral progressive opacity of the embryonic nucleus, pulverulent (fine powderish); zonular/lamellar: affecting only cer-
tain layers between nucleus and cortex.
V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343 337

methods, and by NMR methods, notably by the groups of Gronenborn crystallin has recently been studied [42,43]. Interestingly replacement
[27], Martin [28] and others. of even one of these, namely W43 (by R) is seen to lead to weakening
These studies have revealed several interesting properties of the of the stability, loss of solubility and protein aggregation in human
molecule. For example, γD-crystallin is highly stable, not denatured by γD-crystallin, as in the case of a human patient with cataract [44].
the conventional denaturing agent urea, and needs over 2.5 M of the The two domains are covalently linked by a four/six residue-long
stronger denaturant guanidinum chloride [29]. In contrast, βB1, βA1, connecting peptide which allows them to interact non-covalently
βA3 and βB2 crystallins denature in urea [30,31]. The free energy of through inter-domain contacts. This inter-domain interface is com-
denaturation of human γD-crystallin has been estimated to be about posed of a cluster of six hydrophobic residues and two pairs of polar
8.9 kcal/mol [32], and of γS to be 10.5 ± 0.9 kcal/mol [33]; in compari- peripheral interactions surrounding the hydrophobic cluster. These in-
son, that of αA-crystallin is 6.38 kcal/mol and αB 5.04 kcal/mol [34]. terfacial residues of human γD-crystallin are illustrated in Fig. 2B. The
Similarly, the thermal denaturation temperature of γS is about 74.1 °C hydrophobic cluster consists of M44, F57, and I82 from the N-terminal
and of γD is 83.8 °C [33], although it is pH dependent [35], while domain and V132, L145 and V170 from C-terminal domain. Peripheral
those of β-crystallins are lower; e.g., that of βB2 is around 58 °C [36] pair-wise interactions are between Q55/Q143 and R80/M147. How
and βB1 about 67 °C [37]. It is also interesting to note that the stability such a hydrophobic inter-domain interface contributes to the folding
of the N-terminal domains of γ-D and γ-S crystallins are inherently and stability of the molecule has been described [32].
lower than those of their C-terminal domains [33]. Also, γ-crystallins Quite apart from the aromatics, several other residues (e.g., arginine,
are kinetically stable, making them long-lived. The half-life of the initial cysteine, proline, serine) also contribute to the structure and stability of
unfolding step of γD-crystallin is estimated to be around 19 years [38]! human γ-crystallins. Replacement of even one such ‘critical’ residue by
mutation can lead to drastic changes in the protein's properties, which
3. Roles of some constituent amino acid residues in the structure translate into compromise in transparency of the lens and cataract.
and stability; relation between structural changes and
lens malfunction 4. Section II: mutations in human γ-crystallins and pathology

While it is true that the Greek key motif is adopted by a variety of Table 2 lists as many as 30 such cataract-associated mutations in
proteins with differing primary structural sequences [2], there are human γC, γD and γS crystallins reported to date in human patients.
some constituent residues that play an important role in maintaining (It is interesting to note that most of these are single point mutations,
the architecture of human γ-crystallins. These have been studied in while some others are truncation mutations, one of them is a deletion
great detail by a number of workers cited above. variant, and one a 5-base pair insertion.) They are all inherited (congen-
The role played by the aromatic residues needs special mention. ital) cataracts with no post-translational or age-related modifications in
Human γC-crystallin has a total of 22 aromatic residues (14Y, 4W and phenotype. These make it easy to clone the variant by site-directed mu-
4F), γD has 24 of them (14Y, 4W and 6F), while γS-has 27 of them tagenesis, isolate and purify the molecule and compare its structural
(14Y, 4W and 9F). Of these, the Y residues have a particularly important properties with those of the wild type protein, so as to understand
role in the stability of the Greek key fold, by placing themselves in what what the mutation has done to the properties of the protein.
has been termed as the tyrosine corner [39], wherein a tyrosine residue The role played by the arginine residues (20 of them in human γC,
near the beginning or the end of an antiparallel β-strand hydrogen 21 in γD, and 13 in γS) in maintaining the overall structural folding of
bonds with a residue (often the Y-4 residue, or Δ4 Tyr corner) thus sta- the native protein is exemplified in the Table, where we see that replac-
bilizing the β-barrel structure of the Greek key motif. For example, in ing any of these by other amino acid residues (e.g., R15, R37, R59, R77S
γC-crystallin we find two Δ4 Tyr type of homologous tyrosine corners, and R140X in γD, R48 and R168 in γC) is associated with congenital
one in N-terminal domain around Y63 (Y56, Y63 and W69) and another
one in the C-terminal domain around Y151 (Y144, Y151, W157), and
the same is true with γD, while the tyrosine corners in γS are (Y60,
Y67 and W73) and (Y150, Y157, W163).
Fig. 2A highlights two such tyrosine corners in human γD-crystallin,
those of Y63 and Y151. Yang et al. [40] have recently dissected the con-
tributions of the β-hairpin tyrosine corners to the stability and long life
of the γ-crystallins.
(Note: The amino acid residues in several publications are numbered
counting the N-terminal starting residue methionine as number 1
(e.g., P24T, R77S), while other publications discount met 1 and count
residues as (P23, R76). In order to avoid any confusion and maintain
uniformity, we number residues here counting the starting methionine
as met 1).
Besides tyr, the clustering of aromatic side chains within the interior
of the molecule is a feature that allows the γ-crystallins to take on a
compact shape with all nonpolar residues buried within. These aromatic
residues cluster into several structural elements. Besides the tyrosine
corners, six aromatic pairs (four tyr/tyr, one tyr/phe and one phe/phe)
are present in the β-hairpin sequences of the Greek key. The role of
these aromatic pairing to the stability of the molecule has been
highlighted by Kong and King [41]. The role that trp residues (W) play
is also worth commenting upon. In human γC and γD crystallins, all
the four trp residues are well buried at positions 43, 69, 131, 157. Of
the four trp residues, W69 and W157 are surrounded by aromatic
Fig. 3. Mutation leads to increased surface exposure. Panel A: An overlay of the wild type
amino acids: Y56 and Y63 in the case of W69, and Y144 and Y151 sur- (green; PDB code 2M3T) and S39C mutant (the red arrow points to the S residue). Panel
round W157. Similarly, in γS, the trps are in positions 47, 73, 137 and B: Y11, F16 and S39 residues in the wild type. Panel C: Y11, F16 and C39 in the mutant
163. How these trp and tyr/cys clusters go to stabilize and protect γ- model. Panel D: an overlay of B and C.
338 V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343

cataracts. The γD mutants R37S and R59H are seen to be more prone to change in local environment in the vicinity of residue 39, we built a mo-
crystallization than the wild type protein [65]. lecular model of the C39 mutant using as template the γS-crystallin
And the role played by cysteine residues (6 in γD, 8 in γC and 7 in structure provided by Kingsley et al. [95]. Fig. 3 shows the overlay of
γS) has been studied in detail by the Pandes, who have, for example, the structures of wild type (WT) human γS and the S39C mutant. As
shown how when R15 is replaced by C in γD-, the resultant mutant can be seen in Fig. 3C, the Y11 and, particularly, the F16 side chains
molecule undergoes intermolecular disulfide crosslinking, which have moved outward and are somewhat more exposed in the mutant.
initiates protein aggregation [47]. Giblin et al. [94] have suggested that In order to get a molecular understanding of this cataract, namely,
S-thiolation within the molecule may act to delay the insolubilization how this replacement of S by C affects the structure of the mole-
of the lens proteins. cule, we have used site-directed mutagenesis of the mutant S39C
Though the imino acid proline, both through its ring structure and γS-crystallin molecule, cloned, expressed, purified and studied its
the lack of hydrogen-bonding ability, causes a break in the α-helical solution state properties, and compared them with those of native WT
and β-sheet runs, it seems to play a role in the folding of the molecule, and present some preliminary results here. The materials
Greek key motif. For example, in γD-, replacing the residue P24 with T and methods used for model building, cloning, over expression, isolation
(or other residues such as S or V), as reported in detail by the Pandes and purification of the recombinant wild type and S39C mutant of
[56–60], decreases the solubility of the protein by increasing hydropho- human γS-crystallin and the analysis of their structural and stability fea-
bicity and restricting backbone flexibility, altering the temperature- tures were done as described in our earlier publications [62,78,89,91].
dependent solubility and the phase diagram. The roles of other amino We present the results here.
acids in the sequence, whose replacement is associated with lens mal- Turning to the results, we note from Fig. 4A that the gel elution profile
function, are yet to be understood. We present some results obtained of the WT and the mutant S39C are essentially identical; the two mole-
with one such example involving the serine residue below. cules elute at about the same volume, suggesting that the mutant is
quite soluble and has about the same size as the wild type. (There is a
5. Study of a serine mutant: S39C of human γS-crystallin hint of a higher molecular species eluting a little ahead (around
17.5 ml) of the main band in the mutant, but not in the wild type. We
The role of serine residues in the molecule is yet to be clarified. are currently following this up in order to check whether it suggests mo-
There are as many as 13 S residues in human γC, 17 in γD and 11 in lecular aggregation, e.g., dimerization, in the mutant protein.) Fig. 4B,
γS-crystallin. Yet, interestingly, Table 2 reports mutation in just one of which describes the far-UV circular dichroism (CD) profiles of the two
these so far, namely the single point mutation S39C in human γS- proteins in the far-UV region (250–190 nm), shows that the secondary
crystallin, which is associated with micro-cornea and cataract [76]. structure of the two molecules is about the same. The characteristic neg-
S39 is conserved across all γ-crystallins (it is S35 in γC and γD), and is ative band at 218 nm, positive band around 195 nm and a shoulder
thus thought to be important for the structure. To get an idea of the around 206 nm are indicative of a high degree of β-sheet conformation

Fig. 4. S to C change at 39 does not affect the back bone conformation but slightly alters its tertiary structure. Panel A: Gel elution profile of wild type and mutant human γS-crystallin S39C.
If: fluorescence emission intensity in arbitrary units. λexc: 295 nm; the protein concentrations used were 0.825 mg/mL in 50 mM Tris-Cl (pH 7.3). Panel B: Far-UV CD spectra of wild-type
and mutant human γS-crystallin S39C. θ: elipticity in millidegrees. Protein concentration in each case was 0.150 mg/ml in 10 mM sodium phosphate buffer (pH 7.3). The cell path length
was 2 mm and all spectra were recorded at 37 °C, corrected for background buffer signal and each spectrum is an average of 3 independent runs. Panel C: Near UV CD spectra of wild type
and mutant human γS-crystallin S39C. θ: elipticity in millidegrees. Protein concentration in each case was 0.825 mg/ml in 50 mM Tris-Cl (pH 7.3). The cell path length was 10 mm and the
other conditions of measurement were the same as above. Panel D: Intrinsic fluorescence of wild type and mutant human γS-crystallin S39C. If: fluorescence emission intensity in arbitrary
units. λexc: 295 nm; The protein concentrations used were 10 μM (0.2 mg/mL) in 100 mM sodium phosphate buffer (pH 7.0). The cell path length was 3 mm, excitation and emission slits
2.5 nm and spectra were recorded at 37 °C.
V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343 339

Fig. 5. Panel A: Surface exposure of residues in the wild type and mutant human γS-crystallins S39C, monitored using bis-ANS as the extrinsic probe. λexc: 390 nm, If at 490 nm of the probe
was measured as a function of its increasing concentration. The cell path length was 3 mm, excitation and emission slits 2.5 nm, and spectra were recorded at 37 °C. Each point is an average
of 3 independent runs. Panel B: Aggregation tendencies of the wild type and mutant human γS-crystallin S39C, estimated using Nile Red as the extrinsic probe. λexc: 540 nm, If at 605 nm of
the probe was measured as a function of its increasing concentration. The cell path length was 3 mm, excitation and emission slits 10 nm, and spectra were recorded at 37 °C. Each point is
an average of 3 independent runs.

in the secondary structure of the protein chain, expected of the Greek key units) when bound to S39C than when bound to the WT (10 units),
fold. CD spectra in the near-UV region (340–250 nm), shown in Fig. 4C, suggesting that the mutant molecule has opened up a little more, thus
focusing on the micro-environmental region of the aromatic residues, offering a greater surface for dye binding.
shows a slight difference between the two molecules, essentially in Does a more exposed surface allow the molecule to display a greater
their intensities. But such comparisons are better made and more infor- tendency to make intermolecular interactions and thus aggregate? This
mative, rather than near UV-CD, when one looks at the intrinsic fluores- point can be checked by using the aggregation monitor dye Nile Red.
cence features of the aromatic residues, particularly trp, which has the Fig. 5B compares the binding of this neutral dye Nile Red. Upon binding
highest quantum yield of emission (with the wavelength of excitation to the wild type, it displays a broad band maximum centered around
at 295 nm). When the microenvironment of the aromatic residue is non- 651 nm with an intensity of about 0.65 arbitrary units, while upon bind-
polar, the emission maximum occurs in the 320 nm region, and the emis- ing to S39C, the band blue-shifted by 15 nm to 636 nm, and the intensity
sion quantum yield is rather low. But when the residue experiences a doubled to over 1.3 units; this result too suggests that as the S residue is
more polar environment, its emission band is red-shifted and the inten- replaced by the slightly bigger C, the tendency of the protein to offer a
sity of emission increases. Fig. 4D compares the emission profiles of the somewhat more polar surface and promote intermolecular aggregation
two proteins; we find that while the band maximum is only slightly becomes noticeable.
red-shifted (by 1 nm) in the mutant compared to WT, and the band in- The temperature-dependent unfolding of the two proteins was next
tensity here has increased modestly, by about 10% compared to that of studied by following the relative intensity ratios of fluorescence at
the wild type, suggesting that the local environment around the once 350 nm (representative of the red-shifted band maximum obtained
buried aromatics has become somewhat more exposed. upon total unfolding of the molecule) and at 327 nm (representing
Such changes in the surface polarity are better monitored using dyes the native form). Fig. 6A shows that while the WT molecule starts to un-
such as bis-ANS (4,4′-dianilino-1,1′-binaphthyl-5,5′-disulfonate) [96], fold after 60 °C, with a midpoint at around 74 °C, the mutant unfolds at a
and 9-diethylamino-5H-benzo α-phenoxazine-5-(one) called Nile Red much earlier temperature, well below 70 °C, with the midpoint around
[97] as extrinsic probes which, upon binding to exposed hydrophobic 59 °C, suggesting that the mutant has a lower thermal stability. Indeed,
surfaces display significant changes in their emission maxima and in- the mutant starts scattering light already by about 50 °C during the runs,
tensity. Fig. 5A shows that when 140 μM of bis-ANS is titrated with hampering the accuracy of the experiment. Fig. 6B describes such time-
the proteins, its emission intensity is significantly higher (19 arbitrary dependent increase in light scattering. While the wild type molecule

Fig. 6. Mutant has less thermal stability and is prone to aggregate faster. Panel A: Thermal denaturation of wild type and mutant S39C; λexc: 295 nm; The relative emission intensity of the
350 nm band was compared to that of the 327 nm band and monitored as a function of time. Protein concentration in each case was 0.065 mg/ml in 50 mM Tris-Cl (pH 7.3) and heated in
the range of 30–80 °C. The cell path length was 10 mm, excitation and emission slits 10 nm and spectra were recorded at 37 °C. Sample was allowed for 1 min prior to recording the spectra.
Panel B: Thermal aggregation of wild type and mutant at different temperatures. Light scattering was monitored at 600 nm light for 625 s at different temperatures (43.5, 53.5 and 54.0 °C).
A protein concentration of 0.260 mg/ml in 50 mM Tris buffer was used. The cell path length was 10 mm, excitation and emission slits 5 nm and spectra were recorded at 37 °C.
340 V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343

Fig. 7. Legend: Mutant is less stable towards the chemical denaturant GuHCl and capable of forming an intermediate state during unfolding. Guanidine hydrochloride-induced denatur-
ation of wild type (curve A) and S39C (curve B) γS-crystallin. λexc: 295 nm; the relative emission intensity of the 350 nm band (of the denatured form) was compared to that of the 327 nm
band (of the native protein) and monitored as a function of denaturant concentration. The solid line indicates the fitted data and solid blocks stand for raw data. Protein concentration in
each sample was fixed at 0.2 mg/ml in 50 mM Tris buffer, 1 mM EDTA and 5 mM DTT. The cell path length was 3 mm, excitation and emission slits 5 nm, and spectra were recorded at 37 °C.

remains clear in solution at 54 °C for as long as 600 s, as monitored by interactions between the two residues, with the centroid–centroid dis-
light scattering at 600 nm, the mutant aggregates to produce scattering tance between Y11 and F16 increasing to 6.2 Å in the mutant when
particles even when allowed to stand at 53.5 °C at 400 s; this is further compared to 5.5 Å in the wild type γS-crystallin. Both S39 and C39
hastened to 300 s, when it is incubated at 54 °C. form main chain–main chain H-bonds with the N of Y11. The side
We next studied the chemical denaturant-induced unfolding of the chain oxygen of S39 forms H-bonds with the main chain N atom of
mutant. The denaturant guanidine hydrochloride (GuHCl) was used to F16 and the O of D13. The larger S atom in the mutant forms these
unfold the protein and the relative fluorescence ratio (350 nm/ two H-bonds and, in addition, forms H-bonds with the N atom of N15
327 nm, representative of the ratio of unfolded/native forms of the and the O of P68. The S atom is also involved in the aromatic–sulfur in-
protein) was used as the assay. As Fig. 7A shows, the wild type molecule teraction with the F16, the centroid-sulfur atom distance being 4.5 Å.
displays a clear, sharp two-state transition from the native to the un- However, the larger bulk of the S atom pushes out F16, which also
folded form, with the midpoint of unfolding occurring at the concentra- leads to a disruption, in the mutant. Cation-pi interaction seen between
tion (Cm) of 2.8 M GuHCl. Data from this equilibrium native to the F16 and K41 (distance 3.7 Å) in the wild type protein. Fig. 8 gives a com-
unfolded two-state (N–U) curve were fitted to the two-state transition parison in our model of the inter-atom interactions between Y11, F16
model of Greene and Pace [98], yielding a free energy (ΔG0N–U) value of and S39 (or C39 in the mutant), while Table 4 estimates the distances.
9.145 ± 0.30 kcal/mol. In contrast, the denaturation curve (Fig. 7B) for Note the small but noticeable increase in the distance between the
the mutant displays a multiple step process; the first transition already backbone C atoms of residues 16 (F16) and 39 (S39 or C39) (from
occurs well before 1.5 M GuHCl, and the second at 2.8 M GuHCl, just as 6.60 to 7.34 Å) and also of the N atoms (5.42 to 5.991 Å), while there
with the WT molecule. When these data were analyzed using the three- is some shrinkage between those of Y11 and S/C39.
state model, proposed by Clark et al. [99], we obtained a Cm value of It is also worth pointing out that the mutation involves the replace-
0.73 M GuHCl and a (ΔG0N–I) value of 4.32 ± 0.26 kcal/mol, and a Cm ment of serine (with a relative polarity ranking of 14 (based on the rel-
value of 2.8 M GuHCl for the second transition (I–U) with a ΔG0I–U ative hydrophobicity scale), an average percentage of about 8% burial in
value of 4.88 ± 0.20 kcal/mol (see Table 3). The behavior of S39C is proteins, a van der Waals volume of 73 Å3, and accessible surface area of
thus reminiscent of some other mutants that we had described earlier,
e.g., V42M human γS- and A36P γD-crystallins.
We next examined the changes that the mutation brings about in
our molecular model which was built following protocols described in
our earlier reports [62,78,89,91] using the Swiss-Model workspace
[100] and refined by molecular dynamics procedures in Gromacs 4.5.5
[101]. Fig. 3, shown above, suggests that in the wild type molecule, res-
idues Y11 and F16 are in close contact, and S39 is within interacting dis-
tance. When it is replaced by C in the mutant, both Y11 and F16 move
outward in order to accommodate the bulkier sulfur atom in C39. Fur-
ther analysis shows that there is a change in the aromatic–aromatic

Table 3
Equilibrium unfolding parameters for wild type human γS-crystallin and its S39C mutant.

Protein Equilibrium transition 1 Equilibrium transition 2

Cm Apparent ΔG0 P value Cm Apparent ΔG0 P value

Wild type 2.8 M 9.15 ± 0.3 0.001


[ΔG0N–U]
Fig. 8. Mutation alters the inter-atomic distances: H-bond interactions around residue 39
S39C 0.73 M 4.32 ± 0.26 0.002 2.8 M 4.88 ± 0.20 0.002
in the wild type and S39C mutant of γ-S crystallin. (A) Wild type protein, with the side
[ΔG0N–I] [ΔG0I–U]
chain oxygen of S39 shown as a Corey–Pauling–Koltun rendering and (B) the S39C
[Cm] = transition midpoint in the units of M GuHCl; [ΔG0] = free energy of unfolding in mutant, with the sulfur atom rendered in CPK. Arrows point to hydrogen bonds between
the absence of GuHCl in units of kcal/mol−1; N–U refers to native to unfolded transition, neighboring residues and residue 39. Note that the H-bond pattern between these
N–I native to intermediate and I–U intermediate to unfolded state. residues is not significantly altered in the mutant.
V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343 341

Table 4 who has cataract, PFV and microcornea, has revealed a single point mu-
Alterations in the inter-atom distances upon mutation. tation in human γC crystallin (R48H), but it is not clear what role the
Distance, 2M3T, wild S39C mutant 2M3T, wild S39C mutant mutation plays in each one of these conditions. Work of this kind is ex-
in Å Type C–C, C–C type N–N N–N pected to give us an indication of the role that γ-crystallins play in non-
distance, Å distance, Å distance, Å distance, Å lenticular parts of the eye.
Y11–F16 6.12 6.14 6.57 6.98
Y11–S/C39 5.50 4.82 5.24 4.74
Conflict of interest
F16–S/C39 6.60 7.34 5.42 5.99

C–C: backbone carbon to carbon distance, and N–N backbone nitrogen atoms distance,
None of the authors has any financial disclosures to make.
both in Å units. 2M3T refers to the PDB code for human γS-crystallin.

Acknowledgments
80 Å2 by Cys (relative polarity ranking 7, average percentage burial
value of 3%, van der Waals volume of 86 Å3 and accessible surface area
This work was funded by the Hyderabad Eye Research Foundation.
of 104 Å2) [102]; in other words, even this rather minor alteration
AM thanks the Science Academies of India for a summer research
weakens the stability of the molecule and makes it a little more hydro-
fellowship. We thank Drs. Yogendra Sharma and Rajeev Raman of the
phobic on the surface, causing the tendency to self-aggregate and
Centre for Cellular and Molecular Biology, Hyderabad for the use of
generate light scattering particles which, in the lens, can lead to the clin-
the circular dichroigraph, and for helpful comments. Also, owing to
ical phenotype of lens opacification and cataract. We had earlier shown
the page and reference number limitations, we could not cite all rele-
a similar genotype–molecular phenotype correlation with the muta-
vant original research articles and thereby apologize to the authors
tions V42M and D26G in human γS-crystallin [89,91], and Bharat et al.
whose publications are not cited.
[92] have confirmed this in greater detail in V42M, using NMR spectros-
VPRV and DB conceived the idea, VPRV and AM did much of the ex-
copy; others had shown similar effects with the mutant G18V of the
perimental work, VPRV and DB analyzed and interpreted the data and
same molecule [87,88], and DiMauro et al. [103] have shown that
wrote the manuscript, IK participated in the analysis, modified the man-
even a subtle alteration such as acetylation of gly1 and lys 2 in human
uscript, and checked all the references, SC did the protein modeling
γD-crystallin promotes aggregation of the otherwise highly soluble
work, helped analyze the data, and modified the manuscript. All the au-
molecule. We had earlier argued that the structural integrity of the
thors read and cleared the final manuscript.
Greek key fold seen in lens crystallins appears essential for lens trans-
parency, and a minor distortion in even one of the Greek key folds in
lens crystallins leads to compromise in lens transparency [62]. The pre- References
liminary results we have described above with S39C appear to be in ac- [1] R.C. Augusteyn, On the growth and internal structure of human lens, Exp. Eye Res.
cord with this argument. We still need to investigate the ultimate status 90 (2010) 643–654.
of the C residues in the self-aggregation. [2] A. Mishra, B. Krishnan, S.S. Swaroop, Y. Sharma, Microbial βγ-crystallins, Prog.
Biophys. Mol. Biol. 115 (2014) 42–51.
[3] R.P. Barnwal, M.K. Jobby, K.M. Devi, Y. Sharma, K.V.R. Chary, Solution structure and
6. Section III: gamma crystallins outside the lens calcium binding properties of M-crystallin, a primordial βγ-crystallin from
Archaea, J. Mol. Biol. 386 (2009) 675–689.
[4] E. Serebryany, J.A. King, The βγ-crystallins: native state stability and pathways to
Having looked at the presence of, and the role played by γ- aggregation, Prog. Biophys. Mol. Biol. 115 (2014) 32–41.
crystallins in the lens, we now move to their presence outside the [5] C. Slingsby, G.J. Wistow, A.R. Clark, Evolution of crystallins for a role in the
lens, in other tissues of the eye. While there are many reports for the vertebrate eye lens, Protein Sci. 22 (2013) 367–380.
[6] G.J. Wistow, C. Slingsby, Editorial for special issue: crystallins of the eye, Prog.
non-lenticular presence and functions of α-crystallins, and also some Biophys. Mol. Biol. 115 (2014) 1–2.
members of the β-crystallin family, particularly in the retina, the roles [7] A. Shiels, J.F. Hejtmancik, Genetics of human cataract, Clin. Genet. 84 (2013)
of γ-crystallins are just getting to be understood [104,105]. Based on 120–127.
[8] J.S. Richardson, Beta-sheet topology and the relatedness of proteins, Nature 268
the idea that γS-crystallin is an outlier of the γ family, with some char- (1977) 495–500.
acteristics that appear more likely to retain non-lens function, Sinha [9] H. Zhao, Y. Chen, L. Rezabkova, Z. Wu, G. Wistow, P. Schuck, Solution properties of
et al. [106] looked for the Crygs gene in the mouse eye and found the γ-crystallins: hydration of fish and mammal γ-crystallins, Protein Sci. 23 (2014)
88–99.
γS mRNA to be present in the retina as well as the cornea, though not
[10] Y. Chen, H. Zhao, P. Schuck, G. Wistow, Solution properties of γ-crystallins:
in the iris. They wondered whether γS may have a role as a stress- compact structure and low frictional ratio are conserved properties of diverse
modulator in the tissues of the eye. Most of these studies so far have γ-crystallins, Protein Sci. 23 (2014) 76–87.
concentrated on the murine eye, where they are suggested to play the [11] M. Delaye, A. Tardieu, Short-range order of crystallin proteins accounts for eye lens
transparency, Nature 302 (1983) 415–417.
role of stress-modulators during the development of the eye. Since [12] A. Tardieu, F. Veretout, B. Krop, C. Slingsby, Protein interactions in the calf eye lens:
then, several papers have appeared which have tried to follow up on interactions between beta-crystallins are repulsive whereas in gamma-crystallins
this idea [107,108]. However, all of these studies, while mentioning they are attractive, Eur. Biophys. J. 21 (1992) 1–12.
[13] B. Bax, R. Lapatto, V. Nalini, H. Driessen, P.F. Lindley, D. Mahadevan, T.L. Blundell, C.
βγ-crystallins, focus on the roles of specific β-crystallins such as βA3/ Slingsby, X-ray analysis of beta B2-crystallin and evolution of oligomeric lens
A1 [108], or βB2 [105], and these too on rodent eyes, rather than proteins, Nature 347 (1990) 776–780.
humans. And when talking about γ-crystallins, reference is made invari- [14] R. Jaenicke, C. Slingsby, Lens crystallins and their microbial homologs: structure,
stability, and function, Crit. Rev. Biochem. Mol. Biol. 36 (2001) 435–499.
ably to the mixture rather than a single member such as γS, γD or γC. [15] T. Blundell, P. Lindley, L. Miller, D. Moss, C. Slingsby, I. Tickle, B. Turnell, G. Wistow,
The role of individual members of human γ-crystallins is thus an issue The molecular structure and stability of the eye lens: X-ray analysis of gamma-
worth researching on. This is particularly worthwhile since we find, crystallin II, Nature 289 (1981) 771–777.
[16] G. Wistow, B. Turnell, L. Summers, C. Slingsby, D. Moss, L. Miller, P. Lindley, T.
while working with clinical colleagues in the pediatric ophthalmic sec- Blundell, X-ray analysis of the eye lens protein gamma-II crystallin at 1.9 Å resolu-
tion at our institute, occasional cases of children diagnosed with both tion, J. Mol. Biol. 170 (1983) 175–202.
congenital cataract and persistent fetal vasculature (PFV, wherein the [17] A. Basak, O. Bateman, C. Slingsby, A. Pande, N. Asherie, O. Ogun, G.B. Benedek, J.
Pande, High-resolution X-ray crystal structures of human gamma D crystallin
hyaloid artery involved in nourishing the growth of the eye through
(1.25 Å) and the R58H mutant (1.15 Å) associated with aculeiform cataract, Mol.
vasculo- and angiogenesis does not regress from the anterior portion Biol. 328 (2003) 1137–1147.
of the eye at the appropriate juncture, thus leaving the vasculature to [18] A.G. Purkiss, O.A. Bateman, J.M. Goodfellow, N.H. Lubsen, C. Slingsby, The X-ray
persist and block vision). Zhang et al. [109] had seen the expression of crystal structure of human γS-crystallin C-terminal domain, J. Biol. Chem. 277
(2002) 4199–4208.
γ-crystallins in the hyaloid tissue of an infant with microphthalmia [19] K. Mahendiran, C. Elie, J.C. Nebel, A. Ryan, B.K. Pierscionek, Primary sequence con-
and PFV. Ongoing genetic analysis in one such patient in our center, tribution to the optical function of the eye lens, Sci. Rep. 4 (2014) 5195.
342 V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343

[20] H. Zhao, P.H. Brown, M.T. Magone, P. Schuck, The molecular refractive function of crystallin gene (CRYGD) mutation causes autosomal dominant congenital cerulean
lens γ-crystallins, J. Mol. Biol. 411 (2011) 680–699. cataracts, J. Med. Genet. 40 (2003) 262–267.
[21] M. Hoshino, K. Uesugi, N. Yagi, S. Mohri, J. Regini, B. Pierscionek, Optical properties [51] D.S. Mackay, U.P. Andley, A. Shiels, A missense mutation in the gamma D crystallin
of in situ eye lenses measured with X-ray Talbot interferometry: a novel measure gene (CRYGD) associated with autosomal dominant “coral-like” cataract linked to
of growth processes, PLoS One 6 (2011) e25140. chromosome 2q, Mol. Vis. 10 (2004) 155–162.
[22] B.K. Pierscionek, J.W. Regini, The gradient index lens of the eye: an opto-biological [52] W.Z. Xu, S. Zheng, S.J. Xu, W. Huang, K. Yao, S.Z. Zhang, Autosomal dominant
synchrony, Prog. Retin. Eye Res. 31 (2012) 332–349. coralliform cataract related to a missense mutation of the gamma D-crystallin
[23] T.L. McMeekin, M.L. Groves, N.J. Hipp, Refractive indices of amino acids, proteins gene, Chin Med J (Engl.) 117 (2004) 727–732.
and related substances, in: J. Stekol (Ed.), Amino Acids and Serum Proteins, [53] X. Shentu, K. Yao, W. Xu, S. Zheng, S. Hu, X. Gong, Special fasciculiform cataract
American Chemical Society, Washington DC, 1964. caused by a mutation in the gamma D-crystallin gene, Mol. Vis. 10 (2004)
[24] B.K. Pierscionek, G. Smith, R.C. Augusteyn, The refractive increments of bovine 233–239.
alpha, beta, and gamma-crystallins, Vis. Res. 27 (1987) 1539–1541. [54] A.O. Khan, M.A. Aldahmesh, F.E. Ghadhfan, S. Al-Mesfer, F.S. Alkuraya, Founder
[25] See the King web page at http://web.mit.edu/king-ab/www/publications/pubs_ heterozygous P23T CRYGD mutation associated with cerulean (and coralliform)
cat.htm#Lens (for all relevant papers on their structural studies on γ-crystallin). cataract in 2 Saudi families, Mol. Vis. 15 (2009) 1407–1411.
[26] See the Pande web page at http://www.albany.edu/chemistry/jpande. [55] G. Yang, C. Xiong, S. Li, Y. Wang, J.A. Zhao, A recurrent mutation in CRYGD is asso-
shtml#recentpub (for relevant papers on their structural studies on γ-crystallin). ciated with autosomal dominant congenital coralliform cataract in two unrelated
[27] See the Gronenborn web page at http://www.structbio.pitt.edu/webusers/amg/? Chinese families, Mol. Vis. 17 (2011) 1085–1089.
page_id=21 (for their structural studies on γ-crystallin). [56] A. Pande, O. Annunziata, N. Asherie, O. Ogun, G.B. Benedek, J. Pande, Decrease in
[28] See the Martin web page at http://www.faculty.uci.edu/profile.cfm?faculty_id= protein solubility and cataract formation caused by the Pro23 to Thr mutation in
5277 (for relevant papers on their structural studies on γ-crystallin). human gamma D-crystallin, Biochemistry 44 (2005) 2491–2500.
[29] M.S. Kosinsky-Collins, J. King, In vitro unfolding, refolding, and polymerization of [57] J.J. McManus, A. Lomakin, O. Ogun, A. Pande, M. Basan, J. Pande, G.B. Benedek,
human gammaD crystallin, a protein involved in cataract formation, Protein Sci. Altered phase diagram due to a single point mutation in human gammaD-
12 (2003) 480–490. crystallin, Proc. Natl. Acad. Sci. U. S. A. 104 (2007) 16856–16861.
[30] O.A. Bateman, A. Sarra, S.T. Van Genesan, G. Kappe, N.H. Lubsen, C. Slingsby, The [58] A. Pande, J. Zhang, P.R. Banerjee, S.S. Puttamadappa, A. Shekhtman, J. Pande, NMR
stability of human acidic beta-crystallin oligomers and hetero-oligomers, Exp. study of the cataract-linked P23T mutant of human gammaD-crystallin shows
Eye Res. 77 (2003) 409–422. minor changes in hydrophobic patches that reflect its retrograde solubility,
[31] K.J. Lampi, P.A. Wilmarth, M.R. Murray, L.L. David, Lens β-crystallins: the role of Biochem. Biophys. Res. Commun. 382 (2009) 196–199.
deamidation and related modifications in aging and cataract, Prog. Biophys. Mol. [59] A. Pande, K.S. Ghosh, P.R. Banerjee, J. Pande, Increse in surface hydrophobicity of
Biol. 115 (2014) 21–31. the cataract-associated P23T mutant of human gammaD-crystallin is responsible
[32] S.L. Flaugh, M.S. Kosinski-Collins, J.A. King, Contributions of hydrophobic domain for its dramatically lower, retrograde solubility, Biochemistry 49 (2010)
interface interactions to the folding and stability of human gammaD-crystallin, 6122–6129.
Protein Sci. 14 (2005) 569–581. [60] P.R. Banerjee, S.S. Puttamadappa, A. Pande, A. Shekhtman, J. Pande, In-
[33] I.A. Mills, S.L. Flaugh, M.S. Kosinski-Collins, J.A. King, Folding and stability of the iso- creased hydrophobicity and decreased backbone flexibility explain the
lated Greek key domains of the long-lived human lens proteins gammaD-crystallin lower solubility of a cataract-linked mutant of γD-crystallin, J. Mol. Biol.
and gammaS-crystallin, Protein Sci. 16 (2007) 2427–2444. 412 (2011) 647–659.
[34] T.X. Sun, N.J. Akhtar, J.J. Liang, Thermodynamic stability of human lens recombi- [61] W. Sun, X. Xiao, S. Li, X. Guo, Q. Zhang, Mutation analysis of 12 genes in Chinese
nant alphaA- and alphaB-crystallins, J. Biol. Chem. 274 (1999) 34067–34071. families with congenital cataracts, Mol. Vis. 17 (2011) 2197–2206.
[35] J.W. Wu, M.E. Chen, W.-S. Wen, W.-A. Chen, C.-T. Li, C.K. Chang, C.H. Lo, H.S. Liu, S.S. [62] V.P.R. Vendra, G. Agarwal, S. Chandani, V. Talla, N. Srinivasan, D. Balasubramanian,
Wang, Comparative analysis of human γD-crystallin aggregation under physiolog- Structural integrity of the Greek key motif in βγ-crystallins is vital for central eye
ical and low pH conditions, PLoS One 9 (11) (2014) e112309, http://dx.doi.org/10. lens transparency, PLoS One 8 (2013) e70336.
1371/journal.pone.0112309. [63] L. Wang, X. Chen, Y. Lu, J. Wu, B. Yang, X. Sun, A novel mutation in gamma D-
[36] M. Maiti, M. Kono, B. Chakrabarti, Heat induced changes in the conformation of crystallin associated with autosomal dominant congenital cataract in a Chinese
alpha- and beta-crystallins: unique thermal stability of alpha-crystallin, FEBS family, Mol. Vis. 17 (2011) 804–809.
Lett. 236 (1988) 109–114. [64] S. Kmoch, J. Brynda, B. Asfaw, K. Bezouska, P. Novak, P. Rezácova, L. Ondrova, M.
[37] K.J. Lampi, Y.H. Kim, H.P. Bachinger, B.A. Boswell, R.A. Linder, J.A. Carver, T.R. Filipec, J. Sedlacek, M. Elleder, Link between a novel human gamma D-crystallin
Shearer, L.L. David, D.M. Kapfer, Decreased heat stability and increased chaperone allele and a unique cataract phenotype explained by protein crystallography,
requirement of modified human betaB1-crystallins, Mol. Vis. 8 (2002) 359–366. Hum. Mol. Genet. 9 (2000) 1779–1786.
[38] I.A.R. Mills-Henry, Stability, unfolding, and aggregation of the gamma D and [65] A. Pande, J. Pande, N. Asherie, A. Lomakin, O. Ogun, J. King, Crystal cataracts: human
gamma S human eye lens crystallins(Ph. D. thesis) Department of Biology, genetic cataract caused by protein crystallization, Proc. Natl. Acad. Sci. U. S. A. 98
Massachusetts Institute of Technology, 2007. (2001) 6116–6120.
[39] J.M. Hemmingsen, K.M. Gernert, J.S. Richardson, D.C. Richardson, The tyrosine corner: [66] S.T. Santhiya, G.S. Kumar, P. Sudhakar, N. Gupta, N. Klopp, T. Illig, T. Soker, M. Groth,
a feature of most Greek key beta-barrel proteins, Protein Sci. 3 (1994) 1927–1937. M. Platzer, P.M. Gopinath, J. Graw, Molecular analysis of cataract families in India:
[40] Z. Yang, Z. Xia, T. Huynh, J.A. King, R. Zhou, Dissecting the contributions of new mutations in the CRYBB2 and GJA3 genes and rare polymorphisms, Mol. Vis.
β-hairpin tyrosine pairs to the folding and stability of long lived γD-crystallins, 16 (2010) 1837–1847.
Nanoscale 6 (2014) 1797–1807. [67] A. Santana, M. Waiswol, E.S. Arcieri, Cabral de Vasconcellos JP, M. Barbosa de Melo,
[41] F. Kong, J. King, Contributions of aromatic pairs to the folding and stability of long- Mutation analysis of CRYAA, CRYGC, and CRYGD associated with autosomal dom-
lived human γD-crystallin, Protein Sci. 20 (2011) 513–528. inant congenital cataract in Brazilian families, Mol. Vis. 15 (2009) 793–800.
[42] N. Schafheimer, J. King, Tryptophan cluster protects human γD-crystallin from ul- [68] E. Heon, M. Priston, D.F. Schorderet, G.D. Billingsley, P.O. Girard, N. Lubsen, F.L.
traviolet radiation-induced photoaggregation in vitro, Photochem. Photobiol. 89 Munier, The gamma-crystallins and human cataracts: a puzzle made clearer, Am.
(2013) 1106–1115. J. Hum. Genet. 65 (1999) 1261–1267.
[43] N. Schafheimer, Z. Wang, K. Schey, J. King, Tyrosine/cysteine cluster sensitizing [69] F. Li, S. Wang, C. Gao, S. Liu, B. Zhao, M. Zhang, S. Huang, S. Zhu, X. Ma, Mutation
human γD-crystallin to ultraviolet radiation-induced photoaggregation in vitro, G61C in the CRYGD gene causing autosomal dominant congenital coralliform cat-
Biochemistry 53 (2014) 979–990. aracts, Mol. Vis. 14 (2008) 378–386.
[44] B. Wang, C. Yu, Y.B. Xi, H.C. Cai, J. Wang, S. Zhou, S. Zhou, Y. Wu, Y.B. Yan, X. Ma, L. [70] W. Zhang, H.C. Cai, F.F. Li, Y.B. Xi, X. Ma, Y.B. Yan, The congenital cataract-linked
Xie, A novel CRYGD mutation (p.Trp43Arg) causing autosomal dominant G61C mutation destabilizes gamma D-crystallin and promotes non-native aggre-
congenital cataract in a Chinese family, Hum. Mutat. 32 (2010) E1939–E1947. gation, PLoS One 6 (2011) e20564.
[45] D.A. Stephan, E. Gillanders, D. Vanderveen, D. Freas-Lutz, G. Wistow, A.D. [71] M. Roshan, P.H. Vijaya, G.R. Lavanya, P.K. Shama, S.T. Santhiya, J. Graw, P.M.
Baxevanis, C.M. Robbins, A.V. Auken, M.I. Quesenberry, J.B. Wilson, S.H.H. Juo, Gopinath, K.A. Satyamoorthy, A novel human CRYGD mutation in a juvenile auto-
J.M. Trent, L. Smith, M.J. Brownstein, Progressive juvenile-onset punctate cataracts somal dominant cataract, Mol. Vis. 16 (2010) 887–896.
caused by mutation of the gamma D-crystallin gene, Proc. Natl. Acad. Sci. U. S. A. 96 [72] F. Ji, J. Jung, A.M. Gronenborn, Structural and biochemical characterization of the
(1999) 1008–1012. childhood cataract associated R76S mutant of human gamma D-crystallin, Bio-
[46] F. Gu, R. Li, X.X. Ma, L.S. Shi, S.Z. Huang, X. Ma, A missense mutation in the gamma chemistry 51 (2012) 2588–2596.
D-crystallin gene CRYGD associated with autosomal dominant congenital cataract [73] O.M. Messina-Baas, L.M. Gonzalez-Huerta, S.A. Cuevas-Covarrubias, Two affected
in a Chinese family, Mol. Vis. 12 (2006) 26–31. siblings with nuclear cataract associated with a novel missense mutation in the
[47] A. Pande, D. Gillot, J. Pande, The cataract associated R14C mutant of human gamma CRYGD gene, Mol. Vis. 12 (2006) 995–1000.
D-crystallin shows a variety of intermolecular disulfide cross-links: a Raman spec- [74] P.R. Banerjee, A. Pande, J. Patrosz, G.M. Thurston, J. Pande, Cataract associated mu-
troscopic study, Biochemistry 48 (2009) 4937–4945. tant E107A of human gamma D-crystallin shows increased attraction to alpha-
[48] L.Y. Zhang, B. Gong, J.P. Tong, D.S. Fan, S.W. Chiang, D. Lou, D.S. Lam, G.H. Yam, C.P. crystallin, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 574–579.
Pang, A novel gamma D-crystallin mutation causes mild changes in protein prop- [75] L. Hansen, W. Yao, H. Eiberg, K.W. Kjaer, K. Baggesen, J.F. Hejtmancik, T. Rosenberg,
erties but leads to congenital coralliform cataract, Mol. Vis. 15 (2009) 1521–1529. Genetic heterogeneity in microcornea-cataract: five novel mutations in CRYAA,
[49] L. Hilal, E. Nandrot, M. Belmekki, M. Chefchaouni, S. El Bacha, B. Benazzouz, Y. CRYGD, and GJA8, Invest. Ophthalmol. Vis. Sci. 48 (2007) 3937–3944.
Hajaji, O. Gribouval, J. Dufier, M. Abitbol, A. Berraho, Evidence of clinical and genet- [76] R.R. Devi, W. Yao, P. Vijayalakshmi, Y.V. Sergeev, P. Sundaresan, J.F. Hejtmancik,
ic heterogeneity in autosomal dominant congenital cerulean cataracts, Ophthalmic Crystallin gene mutations in Indian families with inherited pediatric cataract,
Genet. 23 (2002) 199–208. Mol. Vis. 14 (2008) 1157–1170.
[50] E. Nandrot, C. Slingsby, A. Basak, M. Cherif-Chefchaouni, B. Benazzouz, Y. Hajaji, S. [77] S.T. Santhiya, M. Shyam Manohar, D. Rawlley, P. Vijayalakshmi, P.
Boutayeb, O. Gribouval, L. Arbogast, A. Berraho, M. Abitbol, L. Hilal, Gamma-D Namperumalsamy, P.M. Gopinath, J. Loster, J. Graw, Novel mutations in the
V.P.R. Vendra et al. / Biochimica et Biophysica Acta 1860 (2016) 333–343 343

gamma-crystallin genes cause autosomal dominant congenital cataracts, J. Med. local surface hydrophobicity, Biochem. Biophys. Res. Commun. 443 (2014)
Genet. 39 (2002) 352–358. 110–114.
[78] V. Talla, N. Srinivasan, D. Balasubramanian, Visualization of in situ intracellular [93] Z. Yang, Q. Li, S. Zhu, X. Ma, A G57W mutation of CRYGS associated with autosomal
aggregation of two cataract-associated human γ-crystallin mutants: lose a tail, dominant pulverulent cataracts in a Chinese family, Ophthalmic Genet. (2013)
lose transparency, Invest. Ophthalmol. Vis. Sci. 49 (2008) 3483–3490. http://dx.doi.org/10.3109/13816810.2013.865761.
[79] L.Y. Zhang, G.H. Yam, D.S. Fan, P.O. Tam, D.S. Lam, C.P. Pang, A novel deletion [94] F.J. Giblin, L.L. David, P.A. Wilmarth, V.R. Leverenz, M.F. Simpanya, Shotgun prote-
variant of gammaD-crystallin responsible for congenital nuclear cataract, Mol. omic analysis of S-thiolation sites of guinea pig lens nuclear crystallins following
Vis. 13 (2007) 2096–2104. oxidative stress in vivo, Mol. Vis. 19 (2013) 267–280.
[80] Z. Ren, A. Li, B.S. Shastry, T. Padma, R. Ayyagari, M.H. Scott, M.M. Parks, M.I. Kaiser- [95] C.N. Kingsley, W.D. Brubaker, S. Markovic, A. Diehl, A.J. Brindley, H. Oschkinat, R.W.
Kupfer, J.F. Hejtmancik, A 5-base insertion in the gamma C-crystallin gene is Martin, Preferential and specific binding of human αB-crystallin to a cataract-
associated with autosomal dominant variable zonular pulverulent cataract, Hum. related variant of γS-crystallin, Structure 21 (2013) 2221–2227.
Genet. 106 (2000) 531–537. [96] C.G. Rosen, G. Weber, Dimer formation from 1-amino-8-naphthalenesulfonate cat-
[81] K. Yao, C. Jin, N. Zhu, W. Wang, R. Wu, J. Jiang, X. Shentu, A nonsense mutation in alyzed by bovine serum albumin, a new fluorescent molecule with exceptional
CRYGC associated with autosomal dominant congenital nuclear cataract in a binding properties, Biochemistry 8 (1969) 3915–3920.
Chinese family, Mol. Vis. 14 (2008) 1272–1276. [97] M. Sutter, S. Oliveira, N.N. Sanders, B. Lucas, A. Van Hoek, M.A. Hink, A.J. Visser, S.C.
[82] L. Zhang, S. Fu, Y. Ou, T. Zhao, Y. Su, P. Liu, A novel nonsense mutation in CRYGC is De Smedt, W.E. Hennink, W. Jiskoot, Sensitive spectroscopic detection of large and
associated with autosomal dominant congenital nuclear cataracts and denatured protein aggregates in solution by use of fluorescent dye Nile Red, J.
microcornea, Mol. Vis. 15 (2009) 276–282. Fluoresc. 17 (2007) 181–192.
[83] L.M. Gonzalez-Huerta, O.M. Messina-Baas, S.A. Cuevas-Covarrubias, A family with [98] R.F. Greene Jr., C.N. Pace, Urea and guanidine hydrochloride denaturation of ribo-
autosomal dominant primary congenital cataract associated with a CRYGC nuclease, lysozyme, α -chymotrypsin and β-lactoglobulin, J. Biol. Chem. 249
mutation: evidence of clinical heterogeneity, Mol. Vis. 13 (2007) 1333–1338. (1974) 5388–5393.
[84] M. Kumar, T. Agarwal, S. Khokhar, P. Kaur, T.S. Roy, R. Dada, Mutation screening [99] A.C. Clark, J.F. Sinclair, T.O. Baldwin, Folding of bacterial luciferase involves a non-
and genotype phenotype correlation of alpha-crystallin, gamma-crystallin and native heterodimeric intermediate in equilibrium with the native enzyme and
GJA8 gene in congenital cataract, Mol. Vis. 17 (2011) 693–707. the unfolded subunits, J. Biol. Chem. 268 (1993) 10773–10779.
[85] X.Q. Li, H.C. Cai, S.Y. Zhou, J.H. Yand, Y.B. Xi, X.B. Gao, W.J. Zhao, P. Li, G.Y. Zhao, Y. [100] K. Arnold, L. Bordoli, J. Kopp, T. Schwede, The SWISS-MODEL Workspace: a web-
Tong, F.C. Bao, Y. Ma, S. Wang, Y.B. Yan, C.L. Lu, X. Ma, A novel mutation impairing based environment for protein structure homology modeling, Bioinformatics 22
the tertiary structure and stability of human γC-crystallin (CRYGC) leads to cata- (2006) 195–201.
ract formation in humans and zebrafish lens, Hum. Mutat. 33 (2012) 391–401. [101] D. Van Der Spoel, E. Lindahl, B. Hess, G. Groenhof, A.E. Mark, H.J.C. Berendsen,
[86] H. Sun, Z. Ma, Y. Li, B. Liu, Z. Li, X. Ding, Y. Gao, W. Ma, X. Tang, X. Li, Y. Shen, GROMACS: fast, flexible and free, J. Comput. Chem. 26 (2005) 1701–1718.
Gamma-S crystallin gene (CRYGS) mutation causes dominant progressive cortical [102] http://www.proteinsandproteomics.org/content/free/tables_1/table08.pdf&gt.
cataract in humans, J. Med. Genet. 42 (2005) 706–710. [103] M.A. DiMauro, S.K. Nandi, C.T. Raghavan, R.K. Kar, B. Wand, A. Bhunia, R.H. Nagaraj,
[87] Z. Ma, G. Piszczek, P.T. Wingfield, Y.V. Sergeev, J.F. Hejtmancik, The G18V CRYGS A. Biswas, Acetylation of Gly1 and Lys2 promotes aggregation of human γD-
mutation associated with human cataracts increases gamma S-crystallin sensitivity crystallin, Biochemistry 53 (2014) 7269–7282.
to thermal and chemical stress, Biochemistry 48 (2009) 7334–7341. [104] C. Slingsby, G.J. Wistow, Functions of crystallins in and out of lens: roles in elongat-
[88] W.D. Brubaker, R.W. Martin, 1H, 13C, and 15N assignments of wild-type human ed and post-mitotic cells, Prog. Biophys. Mol. Biol. 115 (2014) 52–67.
γS-crystallin and its cataract-related variant γS-G18V, Biomol. NMR Assign. 6 [105] S. Thanos, M.R.R. Bohm, M.M. zu Horste, V. Prokosch-Willing, M. Hennig, D. Bauer,
(2012) 63–67. A. Heiligenhaus, Role of crystallins in ocular neuroprotection and axonal regenera-
[89] S. Karri, R.B. Kasetti, V.P.R. Vendra, S. Chandani, D. Balasubramanian, Structural tion, Prog. Retin. Eye Res. 42 (2014) 145–161.
analysis of the mutant protein D26G of human γS-crystallin associated with [106] D. Sinha, N. Esumi, C. Jaworski, C.A. Kozak, E. Pierce, G. Wistow, Cloning and map-
Coppock cataract, Mol. Vis. 19 (2013) 1231–1237. ping the mouse Crygs gene and non-lens expression of γS-crystallin, Mol. Vis. 4
[90] V. Vanita, J.R. Singh, D. Singh, R. Varon, K. Sperling, Novel mutation in the gamma-S (1998) 8.
crystallin gene causing autosomal dominant cataract, Mol. Vis. 15 (2009) [107] S.E. Jones, C. Jomary, J. Grist, J. Makwana, M.J. Neal, Retinal expression of γS-
476–481. crystallins in the mouse, Invest. Ophthalmol. Vis. Sci. 40 (1999) 3017–3020.
[91] V.P.R. Vendra, S. Chandani, D. Balasubramanian, The mutation V42M distorts the [108] J. Xi, R. Farjo, S. Yoshida, T.S. Kern, A. Swaroop, U.P. Andley, A comprehensive anal-
compact packing of the human gamma-S-crystallin molecule, resulting in congen- ysis of the expression of crystallins in mouse retina, Mol. Vis. 9 (2003) 410–419.
ital cataract, PLoS One 7 (2012) e51401. [109] C. Zhang, P. Gehlbach, C. Gongora, M. Cano, R. Fariss, S. Hose, A. Nath, W.R. Green,
[92] S.V. Bharat, A. Shekhtman, J. Pande, The cataract-associated V41M mutant of M.F. Goldberg, J.S. Zigler Jr., D. Sinha, A potential role for β- and γ-crystallins in the
human γS-crystallin shows specific structural changes that directly enhance vascular remodelling of the eye, Dev. Dyn. 234 (2005) 36–47.

Potrebbero piacerti anche