throbber
U.C. Marx
`f. Klodt
`M. Meyer
`H. Gerlach
`P Rosch
`W -G. Forssmann
`K. Adermann
`
`One peptide, two topologies:
`
`
`structure and interconversion
`
`dynamics of human
`
`uroguanylin isomers
`
`
`
`guanylin; heat-stable enterotoxin; isomerization;
`
`
`
`
`
`solution structure; topological stereoisomer; uroguanylin
`
`
`
`
`
`The peptide hormone uroguanylin stimulates chloride
`
`Key words:
`
`
`
`Authors' affiliations:
`U. C. Marx, Niedersiichsisches Institut fur
`
`
`
`
`
`
`Peptid-Forschung IIPF), Hannover, Germany,
`
`
`and Lehrstuhl fur Struktur und Chemie der
`Abstract:
`
`
`
`
`Biopolymere, Universitiit Bayreuth, Bayreuth,
`Germany.
`
`
`
`
`
`
`
`
`
`
`secretion via activation of intestinal guanylyl cyclase C (GC-C). It is
`
`
`
`[ Klodt, M Meyer, W-G. Forssmann, and K.
`
`
`
`Adermann,
`
`
`Niedersiichsischcs Institut fur
`
`
`
`
`Peptid-Forschung IIPFI, Hannover, Germany.
`
`H. Gerlach,
`
`
`Lehrstuhl fur Organische Chemie 11,
`
`
`
`
`Universitiit Bayreuth, Bayreuth, Germany.
`
`P. Rosch, Lehrstuhl fur Struktur und Chemie
`
`
`
`
`
`der Biopolymere, Universitiit Bayreuth,
`
`
`Bayreuth, Germany.
`
`
`
`
`
`
`
`unambiguous structure-function relationship of the isomers, we
`
`
`
`
`
`determined the solution structure of the separated uroguanylin
`
`
`
`
`
`isoforms using NMR spectroscopy. Both isomers adopt well-defined
`
`
`
`structures that correspond to those of the isomers of the related
`
`
`
`
`
`characterized by two disulfide bonds in a 1-3/2 4 pattern that
`
`
`
`
`
`causes the existence of two topological stereoisomers of which
`
`
`
`
`
`only one induces intracellular cGMP elevation. To obtain an
`
`
`
`
`
`
`
`
`
`peptide guanylin. Furthermore, the structure of the GC-C­
`
`the agonistic
`
`Escherichia
`Niedersiichsisches Institut fur Peptid-Forschung
`
`
`
`
`
`
`with guanylin isomers, the conformational interconversion of
`
`
`
`
`
`coli heat-stable enterotoxin. Compared
`
`
`
`uroguanylin isomers is retarded significantly. As judged from
`
`
`
`
`
`
`
`
`
`activating uroguanylin isomer A closely resembles the structure of
`
`
`
`
`
`
`
`
`
`
`Correspondence to:
`Dr. Knut Adermann
`
`
`
`(!PF)
`Feodor-Lynen-Strasse 31
`
`
`
`
`D-30625 Hannover, Germany
`
`Phone: int +49-511-5466262
`Fax: int +49-511-5466102
`at 37°C with an equilibrium
`
`E-mail: knutadermann@compuserve.com
`
`pH-dependent mutual isomerization
`
`
`
`
`
`isoforms are stable at low temperatures, but are subject to a slow
`
`
`
`
`
`
`
`chromatography and NMR spectroscopy, both uroguanylin
`
`
`
`Ta cite this article:
`
`Marx, U.C., Klodt, J., Meyer, M., Gerlach, H., Rosch,
`
`
`
`
`
`
`
`P., Forssmann, W.-G., Aderrnann, K. (1998) One peptide,
`
`
`
`two topologies: structure and interconversion
`
`
`
`isomers. dynamics of human uroguanylin
`
`f. Peptide
`Res. P, 229-240.
`
`
`
`
`
`magnetic resonance; NOE, nuclear Overhauser effect, also used for
`
`
`
`
`
`
`
`ISSN 1397-002X
`
`
`
`NOESY cross peak; NOESY, NOE spectroscopy; RMSD, root-mean-
`
`229
`
`
`
`isomer ratio of approximately 1: 1 . The conformational exchange is
`
`
`
`
`
`most likely under the sterical control of the carboxy-terminal
`
`
`
`
`
`leucine. These results imply that GC-C is activated by ligands
`
`
`
`
`
`
`
`
`
`
`
`exhibiting the molecular framework corresponding to the structure
`
`
`
`
`
`of uroguanylin isomer A.
`
`Abbreviations:
`
`
`
`cGMP, cyclic 3 ',5 '-guanosine monophosphate;
`
`
`
`Clean-TOCSY, TOCSY with suppression of NOESY-type cross peaks;
`
`
`
`
`
`
`
`
`
`DG, distance geometry; DQF-COSY, double-quantum filtered COSY;
`
`
`
`
`
`
`
`DSS, 2,2-dimethyl-silapentane-5-sulfonic acid; GC-C, guanylyl
`
`
`
`
`
`cyclase C; JR-NOESY, 2D NOESY spectrum acquired with a jump­
`
`
`
`
`
`return observe pulse; MD, molecular dynamics; NMR, nuclear
`
`
`
`
`
`Dates:
`Received 23 January 1998
`
`
`
`Revised 2 March 1998
`
`Accepted 21 March 1998
`
`
`
`
`
`Copyright© Munksgaard 1998
`
`Bausch Health Ireland Exhibit 2010, Page 1 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`U.C. Marx et al . Structure and dynamics of uroguanylin isomers
`
`
`
`square deviation; SA, simulated annea l ing; ST, Escherichia
`coli
`
`
`
`heat-stable enterotoxin; STaR, E. coli heat-stable enterotoxin
`
`amino acids of guanylin and a 22-residue guanylin fragment
`produced by tryptic digestion of the recombinant pro­
`hormone consist of two species each (molar ratio, 1:1) which
`can not be separated (16). One of those isoforms (A) adopts
`a well-defined structure with a right handed spiral confor
`mation similar to that of ST obtained from X-ray crystal
`lography [22). The second isoform (B) adopts a less
`well-defined left-handed spiral conformation, and therefore,
`is suggested to have a lower affinity to the receptor GC C
`(16). This isomerism was confirmed by NMR spectroscopy
`for guanylin containing 17 amino acids (23). Two compo
`nents were observed clearly during high-performance liq­
`uid chromatography (HPLC) analysis for this peptide only
`and the smaller guanylin of 15 residues. Rapid intercon­
`version between the two compounds, however, prevented
`the characterization of the isolated molecular species (15,
`23). Recently, the synthesis of two different topological iso
`mers of human uroguanylin was reported (24). Polyclonal
`antisera against these isoforms were used to detect
`uroguanylin and its corresponding 10-kDa prohormone in
`urine and plasma (11). Although not shown on the structural
`level of the peptides, this finding suggests that both stere
`oisomers of uroguanylin are present in the body. In addi­
`tion to the possible existence of other specific receptors for
`the GC-C-active isomer of uroguanylin, nothing is known
`about the structure and function of the peptide's B form
`that does not cause an increase of intracellular cGMP.
`The purpose of this study was to establish an unambigu­
`ous structure-activity relationship of uroguanylin stereoi­
`somers by combining the NMR-spectroscopical structure
`determination with the GC-C-activating potential and the
`dynamical characteristics of uroguanylin isomers. The re
`sults obtained are a prerequisite to model the interaction
`of ligands with the target protein GC-C with respect to the
`inhibition of GC-C activity and enterotoxic infections.
`
`
`
`Experimental Procedures
`
`
`
`Peptide synthesis, analytical c h romatography and polarimetry
`
`Uroguanylin-16 and uroguanylin-24 were synthesized on a
`preloaded TentaGel-S-PHB-LeuFmoc resin (Rapp Polymere).
`Disulfides of uroguanylin-16 were introduced selectively by
`air oxidation of Cys4 and Cys12 followed by iodine treat­
`ment of acetamidomethyl-protected cysteine residues 7 and
`15. After formation of the second disulfide, two stereoiso­
`mers were obtained and separated by reversed-phase HPLC
`at a temperature of 15 °C. Uroguanylin-24 was prepared cor
`
`
`
`
`
`receptor; TOCSY, total correlation spectroscopy; TPPI, time­
`
`
`
`
`
`
`
`proportional phase incrementation.
`
`Uroguanylin is a small mammalian peptide hormone that
`is known to be involved in the regulation of epithelial wa­
`ter and electrolyte transport. It is related structurally and
`functionally to the agonistic peptide guanylin (1-3). A tar
`get protein for uroguanylin is guanylyl cyclase C (GC C),
`also known as an Escherichia coli heat-stable enterotoxin
`(ST) receptor (STaR) (4, 5 ). Activation of GC-C increases in
`tracellular production of cyclic guanosine monophosphate
`(cGMP) (3, 6), thereby stimulating cGMP dependent protein
`kinase type II (cGKII) [7, 8). cGKII in turn is responsible for
`chloride secretion by the activation of cystic fibrosis trans
`membrane conductance regulator (CFTR), resulting in in­
`testinal fluid and chloride accumulation (9 ).
`Uroguanylin of which different molecular forms were
`identified in urine (3, 6), blood (10, 11), and the intestine (12,
`13) is related closely to guanylin on the level of primary
`structures (Fig. 1). Both peptides are characterized by two
`disulfide bonds in relative positions 1 3 and 2 4, which are
`crucial for biological activity (2, 14, 15 ). The disulfides are
`located within a short sequence of 12 amino acid residues.
`This structural element leads to the existence of two dis­
`tinct topological isoforms of each peptide (15, 16). The en
`terotoxin ST exhibits an equivalent pattern of disulfides,
`but has an additional disulfide loop which may enhance its
`conformational rigidity (17-19 ). In binding assays, both pep­
`tides compete with ST for their common receptor GC C.
`However, the endogenous peptides are less potent activa­
`tors of GC-C than ST is (10, 20, 21).
`The topological stereoisomers of guanylin and uroguany­
`lin exhibit the same cysteine connectivity, but are confor­
`mationally distinct molecules. A synthetic derivative of 13
`
`Uroguan y li n ( u r i n e )
`
`N D DCELCVNVACTGCL
`
`1 �--�
`
`1 6
`
`
`
`U r o gu an y li n ( b l o o d )
`FKTLRTIANDDCELCVNVACTGCL
`
`Guany l i n
`
`PGTCEICAYAACTGC
`
`
`
`E . coli
`�
`
`ent e r o t o x i n ST
`
`NTFYCCELCCNPACTGCY
`
`1. Amino acid sequences and disulfide pattern of GC-C
`activating peptides. The primary structures of human uro
`guanylin and guanylin are shown. ST is from a human strain of
`(41) with the GC C binding and toxic domain located
`between the outer cysteine residues.
`
`Figure
`
`E. coli
`
`2 3 0 I r Peptide
`
`R e s . 5 1 , 1998 I 2 2 9-240
`
`Bausch Health Ireland Exhibit 2010, Page 2 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`respondingly. A detailed synthetic procedure was reported
`recently by Klodt et al. (15 ). The identity of synthetic pep­
`tides was confirmed by electrospray mass spectrometry and
`sequence analysis by Edman degradation. For the cGMP
`T84 cell assay, synthetic peptides were used according to
`the net peptide content as determined by amino acid analy­
`sis. The HPLC study of the mutual conversion of purified
`uroguanylin isoforms was carried out using a C18 column
`(Nucleosil C18 PPN, Macherey & Nagel, 2 x250 mm, 5 µm,
`100 A, buffer A: 0.06% trifluoroacetic acid; buffer B: 0.06%
`trifluoroacetic acid in 80% acetonitrile; flow rate, 0.2 mL/
`min; UV detection at 215 nm; peptide concentration, 1 µg/
`10 µL). Polarimetry was carried out on a Perkin-Elmer 241
`polarimeter. Sample concentration was 1 mg/mL (0.60 mM)
`using 10-cm cells. The optical rotation at 18 °C was detected
`at 5 89 nm (Na), 5 78 nm (Hg), 5 4 6 nm (Hg), 436 nm (Hg)
`and 365 nm (Hg). Starting from pure isoforms the changes
`of the optical rotation were examined for 120 h. For the first
`24 h, the samples were stored at room temperature; from
`24 to 120 h, the samples were incubated at 3 7 °C between
`the measurements.
`
`NMR spectroscopy
`
`Two-dimensional NMR spectra were obtained on commer­
`cial Bruker AMX6oo and AMX400 spectrometers at n°C by
`standard methods (25, 26). Peptide concentrations were: 1.5
`mM, pH 3.9 (uroguanylin-24); 4.9 mM, pH 3.3 (uroguanylin-
`16, isomer A); and 4.1 mM, pH 3.3 (uroguanylin-161 isomer B),
`in H2O/D 2O (9:1, v/v, 500 µL). The H2O resonance was
`presaturated by continuous coherent irradiation at the H2O
`resonance frequency before the reading pulse, except the JR­
`NOESY. The sweep widths in w1 and CO2 were 9 ppm (5434.8
`Hz on AMX6oo and 3 623.2 Hz on AMX400 spectrometer).
`Quadrature detection was used in both dimensions with the
`time proportional phase incrementation (TPPI) technique in
`co,. In CO2 4000 data points were collected, and 512 data points
`in co1 • Zero filling to 1000 data points was used in w1. Spec­
`tra were multiplied with a squared sinebell function phase
`shifted by rr,/41 rc/3, and rc/2 for the NOESY spectra, by rc/4
`for the Clean-TOCSY and the DQF-COSY spectra. Baseline
`and phase correction of the sixth order was used. Data were
`evaluated on X-Window workstations with the NDee pro­
`gram package (Software Symbiose, Bayreuth). For the se­
`quence-specific assignment of spin systems and the
`evaluation of the NOESY distance constraints for the pep­
`tides, data from the following 600 MHz spectra were used:
`DQF-COSY, Clean-TOCSY with a mixing time of So msec,
`NOESY with a mixing time of 200 msec and JR-NOESY with
`
`U.C. Marx et al . Structure and dynamics of uroguanylin isomers
`
`a mixing time of 200 msec. For estimation of the coupling
`constants, DQF-COSY spectra were recorded with 8000 data
`points in w2 and 512 data points in co, and were processed
`without window function. Coupling constants were mea
`sured between the antiphase peaks of the resonances using
`a Lorentzian function for peak fitting. The time dependent
`10 NMR spectra were obtained from the AMX400 spectrom­
`eter (400 MHz) at 2 5 °C. For the first 24 h, samples were
`stored at room temperature between the measurements;
`from 24 to 92 h, samples were stored at 3 7 °C.
`
`Structure calculations and analysis
`
`The total number of nontrivial unambiguous NOESY cross
`peaks used for structure calculation was 84 for the A form
`of uroguanylin-16 and 69 for the B form (Table 1). These
`cross peaks were divided into three groups according to
`their relative intensities: strong, 0.2 to 0.3 nm; medium,
`0.2 to 0-4 nm; weak, 0.2 to 0.5 nm. In a pseudoatom ap­
`proach, 0.05 nm was added to the upper distance limit for
`distances involving unresolved methyl or methylene pro­
`ton resonances. Also included by an iterative strategy were
`s <\> and 3 x' dihedral angle restraints for isomer A and 4 (\l
`and 2 x' dihedral angle restraints for the B form. Deviations
`of 30° from the calculated angles were allowed in the cal-
`
`
`t..,%�,� '$ • Jf�l�f9)t � �ntrH.>�(h:.<�S.: ���� $:#t�Ni ft<.'>�'N :S�«��t���
`
`
`
`
`
`
`t��t��:�i!!�t�:�$���,�!.!:!,;t����tt�}���!l�\t��,�t�!����ttt��tt�,,��"�����-
`
`� f·f ·:�: {� ,
` : �j{:��··:;�
`H··] t· :-.�-':f
`�t·n>, . :�
`,f
`:: g .
`
`�:-,:_,�$
`'-�-=�
`�:}.�;�
`� . t:$
`
`�4:ii
`:f{��
`·-t�%
`�t���
`�t�ii
`
`t:�t?. �:..:. B)>�����
`�... ��� f: tfr>���
`)�;.�}�'
`:· �?i:�>-.l�:·S�}�
`1Ji /·��···����1
`
`
`�}.:£f)�f'
`
`-:\;< :) �;:
`�f���tr
`
`~5/<� ���""" """ �"""�,� �� A*<"""* ��- �_ .,�
`
`����i�i�+i��+�fii+������~.�-
`
`. ,.,. ••• i�t
`
`::�:t{������
`Jh� ��}l:��$ ��h��t�}��t���J�f
`
`
`··;f:::=�t�::�\����::t ::_.:_����:t�*�����
`
`-��-��=-��,����������:j:::::::::::4.� ..:::.,.,-----,--------,_,,.-
`
`f. Peptide
`Res. ;,, 1998 / 229-240 I 2 3 l
`
`Bausch Health Ireland Exhibit 2010, Page 3 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`A
`
`41
`Glyl
`
`Vall O
`
`I
`
`Thr1 3
`Gius
`�la!
`
`I
`
`I
`
`Leu6
`
`3.8
`ijva
`18
`4.0
`I I
`
`4.2 [
`
`4.4 8
`
`eul 6
`C)'
`
`Asp3
`
`�
`
`C s l 2
`
`4.6
`
`Cysl 5
`
`9
`Asn
`
`4.8
`
`'
`
`'
`
`As
`
`I p2
`-
`
`
`
`
`
`isomers U.C. Marx et al . Structure and dynamics of uroguanylin
`
`culation without penalty. Structure calculations were per­
`formed using a modified ab mitio SA protocol with an ex
`tended version of X-PLOR 3.1 program package (27). The
`structure calculation included floating assignments of
`prochiral groups (28) and a reduced presentation for
`nonbonded interactions for part of the calculation (29) to
`increase efficiency. The calculation strategy is similar to
`those described previously (30, 31). Structure parameters
`were extracted from the standard files parallhdg.pro and
`topallhdg.pro of X-PLOR V3.840 (27). The disulfide bonds
`were included explicitly. For each fragment, 30 structures
`were calculated and 10 structures for each isoform were se­
`lected with the criterion of the lowest overall energy for
`further characterization. SYBYL 6.o (TRIPOS), RASMOL V.
`2.6 (32) and MOLSCRIPT (33) were used for visualization
`of the structure data. To elucidate the stability of the struc­
`tures, we calculated local RMSD using SYBYL 6.o as well
`as XPLOR 3.1. The geometry of the structures was analyzed
`using PROCHECK (34)1 PROMOTIF (35) and XPLOR 3.1.
`
`cG MP T84 cel l assay
`
`Synthetic peptides were tested to assess the specific CC­
`C-stimulating potency as described previously (e.g. 10, 21).
`T84 cells were preincubated with 1 mM isobutylmethyl­
`xanthine for 5 min. Then, peptides were added to the me
`dium in a concentration range of 10-9 to 10-6 M and cells
`were incubated for 60 min. Incubation was stopped by re­
`moval of medium and addition of ice cold ethanol. The
`amount of intracellular cGMP induced by the peptides was
`determined using a specific radioimmunoassay (36).
`
`
`
`Results and Discussion
`
`The isomers of uroguanylin-16 and the amino-terminally
`extended uroguanylin-24 were synthesized and separated
`using standard chromatography techniques as described
`under "Experimental Procedures". The GC-C-activating
`and earlier eluting compound during HPLC is referred to
`as isomer A.
`
`Chemical shift ana lysis
`
`With very little resonance overlap in the 2D NMR spectra
`of the separated isoforms of uroguanylin-16 (Fig. 2), the se
`quence-specific assignment of each species could be per­
`formed by standard methods (26, Table 2). The amide
`proton chemical shifts of Cys7, Val8, Cys12 and Thn3 dif-
`2 3 2 I /. Peptide
`
`Res. 52, ,998 / 2 29-240
`
`9.0 8.8 8.6 8 .4 8.2 8.0 7.8 7.6
`ro2 (ppm)
`
`B
`
`Gly l 4
`
`Vall o
`Valf
`
`' ,
`
`0
`
`" ll
`
`unp
`
`A.la '1
`
`1 Lcµ6
`j I eul 6
`
`0
`
`Thrl3
`
`3 . 8
`
`4.0
`
`4.2 [
`
`4.4 8
`
`•
`
`Cys t
`
`1Cysl 5
`
`As p2
`
`1L ,.
`
`Cys4
`
`C s
`7
`Asn9 y
`
`4.6
`
`4.8
`
`�
`
`9.0 8 . 8 8.6 8.4 8.2 8.0 7.8 7.6
`co 2 (ppm)
`
`Figure 2 . NOESY spectra ( 200 msec) of uroguanylin 16 isomers.
`
`
`
`
`
`
`Fingerprint region of (A) isomer A and (BJ of isomer B. The
`
`intraresidual cross peaks between the amideand Ca-protons of
`
`
`each amino acid are labeled and the chain tracing along the
`
`sequence is indicated.
`
`fer by up to 1 ppm between the two isoforms, indicating
`that the backbone folds in these regions are different (Fig.
`3A). The differences in the chemical shifts of the Ca-pro
`ton resonances with a maximum of 0.3 ppm are less sig­
`nificant, but obvious for the residues Cys7 and Cys12 (Fig.
`3B). Amide proton chemical shifts of the first six and Ca.­
`proton chemical shifts of the first three residues are virtu­
`ally identical for both isoforms, indicating that the
`backbone folds of the amino-terminus of both isoforms are
`similar. In the 2D NMR spectra of a 1:1 mixture of both iso
`mers of the NH 1-terminally extended uroguanylin-24 (data
`
`Bausch Health Ireland Exhibit 2010, Page 4 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`U.C. Marx et al . Structure and dynamics of uroguanylin isomers
`
`,twJ�
`�>*·\�
`
`not shown) only one set of spin systems was found for the
`residues Phe1 to Aspn, indicating that the u amino-termi­
`nal amino acids have the same conformation, most prob­
`ably unstructured or extended, as judged from their
`Ca-proton chemical shifts. For the 13 carboxy terminal
`amino acids, an unambiguous and independent assignment
`of resonances of the A and B forms was possible because
`of clear differences in the chemical shifts of both isoforms.
`The backbone chemical shift values of the last 13 amino
`acids comprising the Cys-rich region of uroguanylin-16 and
`-24 differ by less than 0 . 0 8 ppm for the corresponding
`isoforms with one exception, indicating that the isoforms
`of uroguanylin 24 most probably have a tertiary fold which
`closely resembles the corresponding isomers of uro­
`guanylin-16. Thus, the additional 8 amino acids at the
`amino terminus of uroguanylin-24 apparently do not influ­
`ence the global fold of the backbones of the two isomers.
`
`
`
`
`
`tocol (27) was used with explicit inclusion of disulfide
`bonds. None of the ten selected structures shows NOE vio
`lations more than 0.03 nm, and no structure has angle vio­
`lations more than 5 °. Although only a few nonsequential
`NOEs were assigned, and no typical NOEs and no consecu­
`tive 3 f NHcxH coupling constants for any regular secondary
`structure element were found, the NMR data were suffi­
`cient to define well structured global folds for both
`isoforms. The well-defined global fold of the region from
`Cys4 to Cys15 is reflected by the low backbone root-mean
`square deviation (RMSD) of 0.07 5 nm for the A form and
`0.063 nm for the B form (Fig. 4A). The structures of the iso
`mers A and B of uroguanylin are rather different: the back­
`bone RMSD between the average structures of the A and B
`forms from residue Cys4 to Cys15 is 0.46 nm. For each frag­
`ment, the ten final structures have been deposited in the
`Brookhaven Protein Databank (accession numbers 1UYA
`and 1UYB).
`The structure of the A form of uroguanylin-16 contains
`Structure calculation and analysis of uroguanylin-16 isomers
`three loops, Cys4 to Cys7, Cys7 to Cys12 and Cys12 to
`Cys15 . These loops are arranged in a right-handed spiral
`which is stabilized by disulfide bonds Cys4-Cys12 and Cys7
`Cys15 (Fig. 4B, C). A view along the spiral axis shows that
`the connected cysteines are in line after one complete spi­
`ral turn. The three amino-terminal amino acids are unstruc­
`tured. No elements of regular secondary structure were
`
`
`
`Eighty-four and sixty-nine distance restraints from the 200
`msec-NOESY spectra at 11 °C were collected for uro­
`guanylin-16 isomers A and B and were used together with
`8 and 6 dihedral restraints, respectively, in restrained MD
`calculations (Table 1). For structure calculations of both
`isoforms, a modified ab initio simulated annealing ISA) pro-
`
`
`
`/. Peptide Res. s>, 1998 / 229-240
`
`I 2 3 3
`
`Bausch Health Ireland Exhibit 2010, Page 5 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`U.C. Marx et al . Structure and dynamics of uroguanylin isomers
`
`A
`
`"' 0.5
`•
`z
`
`0.0
`
`· -0.5
`z
`
`B
`
`0.2
`
`0.1
`
`"
`
`-0.1
`
`-0.2
`
`::r:
`tl
`IQ -0.3
`
`Sequence position
`
`2 4
`
`10 1 2 14
`8
`6
`Sequence position
`
`1 6
`
`Figure 3. Chemical shift differences of the backbone protons of
`uroguanylin-16 isomers. (A) Differences of the chemical shifts of
`the backbone amide protons of isomers A and B, (B) differences of
`the chemical shifts of Ca-protons of isomers A and B . The values
`were taken from Table 2.
`
`found for the ten calculated conformations except �- and
`-y-turns. For 80% of the A form structures, the three loops
`are characterized by type IV � turns having low RMSD val­
`ues within the structure family. For five structures, an in­
`verse -y-turn from Asn9 to Ala11 also was found. The
`structure of the 13 carboxy-terminal amino acids of the B
`form can be depicted as a distorted left-handed spiral (Fig.
`4B, C). Where Cys4 and Cys7 are superimposed with their
`corresponding half-cystines Cys12 and Cys1 5 , the loop be­
`tween Cys4 and Cys7 is parallel to the backbone segment
`from Cys12 to Cys7. As for the A form, the three amino
`terminal amino acids are unstructured and no regular sec­
`ondary structure elements except reverse turns were
`identified for the 10 final structures of the B form. All 1 0
`structures show type IV �-turns between Cys4 and Cys7
`and between Cys12 and Cys1 5 with comparable backbone
`angles within the structure family.
`With a backbone RMSD of 0-46 nm for residues Cys4
`to Cys15 between the average structures of A and B forms,
`
`
`
`234 I T- Peptide Res. S l, 1998 / 229-240
`
`the backbone folds of the two isomers differ significantly.
`To identify the segments with the highest degree of differ­
`ences in backbone fold between the two isomers, we cal­
`culated the local RMSD values between the two average
`structures using a five-amino acid window ( 3 7, Fig. 5 ) . The
`RMSD values around Cys4 and the loop around Asn9 and
`Vaho are reduced, indicating that the local structures of the
`A and B forms of uroguanylin-16 resemble each other in
`these sequence regions. From Leu6 to Val8 and from Alan
`to Thn3 the local RMSD is increased, indicating that the
`backbone folds of the A and B forms differ most signifi­
`cantly within this region. The two segments of higher
`RMSD comprise the residues Cys7 and Cys12, for which sig­
`nificant differences of the (j> and 'l' angles between the two
`isoforms also were found. Thus, the local backbone folds
`within the three loops are less different for the two
`isoforms, but their arrangement determines the overall
`backbone fold, mostly restricted by the backbone direction
`around Cys7 and Cys12. This feature also is reflected by the
`high differences in the chemical shifts of these residues be­
`tween the A and B forms (Fig. 3 ) .
`
`Structure comparison with guanyli n a n d E . coli enterotoxin ST
`
`Uroguanylin shows a high sequence homology with the
`related peptide guanylin (Fig. 1) and has a similar ability to
`stimulate GC-C (3, 1 5 ) . The chemical shifts of the backbone
`protons of the cysteines and the following residues of
`uroguanylin resemble those of the guanylin containing 13
`
`amino acids found by Skelton et al. (16 ) and guanylin con­
`
`taining 17 amino acids (23). This phenomenon is most strik­
`mg for Cys7 and Cys12 and their consecutive residues. The
`backbone RMSD values for the cysteine rich region be­
`tween the average structure of uroguanylin-16 isoforms and
`the structures of guanylin-13 isoforms (Brookhaven Protein
`Databank, 1GNA and 1GNB) is 0.14 nm for the A forms and
`0.15 nm for the corresponding B forms. In contrast, the
`RMSD values between the average s tructures of the
`noncorresponding isoforms are higher than 0.4 nm. Al­
`though the guanylin isomers could be characterized only
`in a mixture, this strong conformity of the structures now
`renders it all but certain that guanylin isomer A is actu­
`ally the GC-C-stimulating isomer.
`A comparison of the average structures of uroguanylin-
`16 isomers with the crystal structure of the toxic domain
`of the heat-stable enterotoxin ST ( Fig. 11 (221 PDB acces­
`sion number 1ETN) shows that the A form of uroguanylin-
`16 closely resembles the structure of ST; the backbone
`RMSD for C y s 4 to C y s 1 5 ( numbering according t o
`
`Bausch Health Ireland Exhibit 2010, Page 6 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`
`
`U.C. Marx et al . Structure and dynamics of uroguanylin isomers
`
`
`
`
`
`Isomer A
`
`lsomer B
`
`Cys4
`
`A
`
`4. (A) Best-fit superposition from Cys4 to
`Figure
`Cys15 of the backbone atoms of the 10 final
`structures of each uroguanylin-16 isomer. (left)
`isomer A; (right) isomer B. (BJ Lowest energy
`solution structures of uroguanylin 16 isomers.
`(left) isomer A; (right) isomer B. The lower
`views were obtained from a 90° rotation about a
`horizontal axis. (C) Schematic view of the
`backbone folds and the disulfide connectivities
`of uroguanylin-16 isoforms. (left) isomer A;
`(right) isomer B. Cysteine residues are repre­
`sented by encircled numbers.
`
`8
`
`COOH
`
`�
`
`C 11 ---t-t
`
`uroguanylin-16 ) between the A form of uroguanylin and
`ST is 0.11 nm, whereas the RMSD value for the B form is
`0.45 nm. The known higher activation potency of ST may
`be related to the additional disulfide bond which causes
`a higher rigidity of its three-dimensional structure and,
`thus, a possibly more efficient interaction with the recep­
`tor. Structure calculations of uroguanylin-16 with an ad­
`ditional distance restraint between protons that occupy
`the positions of fictitious sulfur atoms of a third disulfide
`bridge between residues 3 and 8 show that a third disul­
`fide bridge is possible for the A form structure without
`distortion of the peptide backbone. The same calculation
`for the B form resulted in a higher overall energy of these
`structures and a slight violation of the additional fictitious
`
`distance restraint, indicating that a third disulfide bridge
`for the B form structure could be possible but needs a
`higher distortion of the peptide backbone ( data not shown).
`A third disulfide bond apparently would lead to a prefer­
`ence of a structure similar to the A form isomer that was
`found for ST ( 2 2 ) .
`Because GC C i s not activated by uroguanylin isomer B,
`the rigid framework provided by the structures of the A
`form isomers of uroguanylin and guanylin as well as of ST
`is necessary for the binding of these ligands to the extra­
`cellular domain of GC-C and for the activation of the in
`tracellular catalytic domain. The well-defined topology of
`A-type isomers, however, does not seem to be sufficient to
`activate the receptor; this was demonstrated by the chi-
`
`
`
`/. Peptide Res 52, 1998 / 229-140 I 2 3 5
`
`Bausch Health Ireland Exhibit 2010, Page 7 of 12
`Mylan v. Bausch Health Ireland - IPR2022-00722
`
`

`

`U.C. Marx et al . Structure and dynamics of uroguanylin isomers
`
`0.15
`
`0.10
`
`0.05
`
`2
`
`4
`
`6
`
`8 1 0
`Sequence position
`
`12 14
`
`Figure 5 . Local RMSD values between the A and B forms of
`uroguanylin 16 calculated with a five amino acid window (37).
`• heavy atoms; ■
`RMSD, backbone.
`
`meric peptide EDPGTCEICVNVACTGC investigated in
`our laboratory (15 ). Here, the central residues AYA of
`guanylin were replaced by VNV of uroguanylin. Showing a
`guanylin-like isomerism, this derivative generated two iso­
`mers, as detected by HPLC and 2D NMR spectroscopy, but
`had an unexpectedly poor potency to stimulate GC-C in a
`minimum concentration of 10-
`M. To understand the indi­
`6
`vidual contribution of single amino acids for receptor bind
`ing and activation, further experiments using systematic
`substitution of amino acids contained in uroguanylin and
`guanylin are necessary.
`
`
`
`
`
`
`
`Stability and conversion of uroguanylin isomers
`
`Because of the known interconversion equilibrium of
`guanylin isomers, the stability of the separated uro­
`guanylin isomers was investigated regarding the influence
`of temperature, pH and solvent. The effect of temperature
`on the stability of uroguanylin 16 isomers A and B in aque­
`ous solution at pH 4.5 was determined by HPLC after dif­
`ferent periods of incubation up to 7 days . The results
`obtained show that both isomers pass through a tempera­
`ture-dependent conversion generating the corresponding
`stereoisomeric peptide without detectable decomposition
`or disulfide exchange at this pH (Fig. 6A, B). Integration
`of HPLC peaks obtained for isomers A and B showed that
`both isomers of uroguanylin-16 are completely stable at
`0°C, whereas a temperature of 60 °C caused an accelerated
`formation of the complementary isomer within about 4
`h. At 3 7 °C, 2 5 % of both uroguanylin-16 isomers are inter­
`converted within 24 h. This and conversion experiments
`carried out at higher temperatures resulted in a mixture
`containing uroguanylin isomers A and B in an approxi-
`
`r Peptide Res. 52, 1998 / 229-240
`2 3 6 I
`
`mately 1:1 ratio with a slight preference toward isomer A.
`It was not possible to shift this equilibrium ratio. Thus,
`isomer A is the thermodynamically slightly preferred mo­
`lecular species of uroguanylin. From the chromatographi­
`cally detected conversion, it is clear that once one of the
`uroguanylin isomers has turned into its stereoisomer, it
`is subjected to a dynamic equilibrium reconstituting the
`original isomer and vice versa. No conversion of the lyo­
`philized uroguanylin isomers after long term storage of up
`to 3 months at -20 °C was observed. Further experiments
`at 3 7 °C clearly show a significant effect of the medium's
`acidity on the transition kinetics between the stereoiso­
`meric forms of uroguanylin containing 16 and 24 amino
`acids (Fig. 6C). Although a decrease of pH below 4.5 did
`not change the kinetics of interconversion, a slight alka­
`line pH adjusted either by ammonium hydrogen carbon­
`ate or by the medium used for cGMP T84 cell assay
`drastically reduced the rate of interconversion. Therefore,
`in principle, it is possible that, under conditions reflect
`ing the pH of blood and within the wide range of mucosal
`acidity (21), both uroguanylin isomers may exist without
`generating a significant amount of their corresponding ste­
`reoisomer. This feature clearly distinguishes uroguanylin
`from the related peptide guanylin whose isoforms have
`been shown to interconvert much more rapidly (15, 16).
`The slow development of the equilibrium b etween
`uroguanylin isomers at alkaline pH indicates that the ion
`ization state of the isomeric molecules strongly influences
`the kinetics of transition between the isome1s of
`uroguanylin 16 and uroguanylin 24. Thus, the terminal
`carboxyl, ionizable side-chains of Asp2, Asp3 and Glu5,
`or those groups able to form intrachain hydrogen bonds,
`may be involved in the control of stabilization of the two
`isomers. After 3 days at alkaline pH, both isoforms decom­
`posed because of disulfide exchange. Comparison of the
`conversion of uroguanylin 16 isomers with the isomers of
`the N-terminally extended uroguanylin-24 resulted in
`identical kinetics for isoforms A and B at a pH of 4.5 and
`7. 7 for both peptides (Fig. 6C). This result clearly demon­
`strates that the isomerization is not affected by the amino­
`terminal region of uroguanylin. Corresponding to the
`equilibrium isomer ratio with a slightly preferred isomer
`A, B-type isomers of both peptides are converted some­
`what faster than A-type isomers. The interconversion of
`uroguanylin 16 isomers also was studied by recording the
`time-dependent change of the o

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket