throbber
NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.
`Lodish H, Berk A, Zipursky SL, et al. Molecular Cell Biology. 4th edition. New York: W. H. Freeman; 2000.
`
`Section 3.1 Hierarchical Structure of Proteins
`
`Proteins are designed to bind every conceivable molecule—from simple ions to large complex
`molecules like fats, sugars, nucleic acids, and other proteins. They catalyze an extraordinary
`range of chemical reactions, provide structural rigidity to the cell, control flow of material
`through membranes, regulate the concentrations of metabolites, act as sensors and switches,
`cause motion, and control gene function. The three-dimensional structures of proteins have
`evolved to carry out these functions efficiently and under precise control. The spatial
`organization of proteins, their shape in three dimensions, is a key to understanding how they
`work.
`
`One of the major areas of biological research today is how proteins, constructed from only 20
`different amino acids, carry out the incredible array of diverse tasks that they do. Unlike the
`intricate branched structure of carbohydrates, proteins are single, unbranched chains of amino
`acid monomers. The unique shape of proteins arises from noncovalent interactions between
`regions in the linear sequence of amino acids. Only when a protein is in its correct three-
`dimensional structure, or conformation, is it able to function efficiently. A key concept in
`understanding how proteins work is that function is derived from three-dimensional structure,
`and three-dimensional structure is specified by amino acid sequence.
`
`The Amino Acids Composing Proteins Differ Only in Their Side Chains
`Amino acids are the monomeric building blocks of proteins. The α carbon atom (C ) of amino

`acids, which is adjacent to the carboxyl group, is bonded to four different chemical groups: an
`amino (NH ) group, a carboxyl (COOH) group, a hydrogen (H) atom, and one variable group,
`2
`called a side chain or R group (Figure 3-1). All 20 different amino acids have this same general
`structure, but their side-chain groups vary in size, shape, charge, hydrophobicity, and reactivity.
`
`Figure 3-1
`
`Amino acids, the monomeric units that link together to form
`proteins, have a common structure. The α carbon atom
`(green) of each amino acid is bonded to four different
`chemical groups and thus is asymmetric. The side chain, or R
`group (red), is (more...)
`
`The amino acids can be considered the alphabet in which linear proteins are “written.” Students
`of biology must be familiar with the special properties of each letter of this alphabet, which are
`determined by the side chain. Amino acids can be classified into a few distinct categories based
`primarily on their solubility in water, which is influenced by the polarity of their side chains
`(Figure 3-2). Amino acids with polar side groups tend to be on the surface of proteins; by
`interacting with water, they make proteins soluble in aqueous solutions. In contrast, amino acids
`with nonpolar side groups avoid water and aggregate to form the waterinsoluble core of proteins.
`The polarity of amino acid side chains thus is one of the forces responsible for shaping the final
`
`MYLAN INST. EXHIBIT 1091 PAGE 1
`
`MYLAN INST. EXHIBIT 1091 PAGE 1
`
`

`

`three-dimensional structure of proteins.
`
`Figure 3-2
`
`The structures of the 20 common amino acids grouped into
`three categories: hydrophilic, hydrophobic, and special amino
`acids. The side chain determines the characteristic properties
`of each amino acid. Shown are the zwitterion forms, which
`exist at the (more...)
`
`Hydrophilic, or water-soluble, amino acids have ionized or polar side chains. At neutral pH,
`arginine and lysine are positively charged; aspartic acid and glutamic acid are negatively
`charged and exist as aspartate and glutamate. These four amino acids are the prime contributors
`to the overall charge of a protein. A fifth amino acid, histidine, has an imidazole side chain,
`which has a pK of 6.8, the pH of the cytoplasm. As a result, small shifts of cellular pH will
`a
`change the charge of histidine side chains:
`
`The activities of many proteins are modulated by pH through protonation of histidine side
`chains. Asparagine and glutamine are uncharged but have polar amide groups with extensive
`hydrogen-bonding capacities. Similarly, serine and threonine are uncharged but have polar
`hydroxyl groups, which also participate in hydrogen bonds with other polar molecules. Because
`the charged and polar amino acids are hydrophilic, they are usually found at the surface of a
`water-soluble protein, where they not only contribute to the solubility of the protein in water but
`also form binding sites for charged molecules.
`
`Hydrophobic amino acids have aliphatic side chains, which are insoluble or only slightly soluble
`in water. The side chains of alanine, valine, leucine, isoleucine, and methionine consist entirely
`of hydrocarbons, except for the sulfur atom in methionine, and all are nonpolar. Phenylalanine,
`tyrosine, and tryptophan have large bulky aromatic side groups. As explained in Chapter 2,
`hydrophobic molecules avoid water by coalescing into an oily or waxy droplet. The same forces
`cause hydrophobic amino acids to pack in the interior of proteins, away from the aqueous
`environment. Later in this chapter, we will see in detail how hydrophobic residues line the
`surface of membrane proteins that reside in the hydrophobic environment of the lipid bilayer.
`
`Lastly, cysteine, glycine, and proline exhibit special roles in proteins because of the unique
`properties of their side chains. The side chain of cysteine contains a reactive sulfhydryl group
`(—SH), which can oxidize to form a disulfide bond (—S—S—) to a second cysteine:
`
`MYLAN INST. EXHIBIT 1091 PAGE 2
`
`MYLAN INST. EXHIBIT 1091 PAGE 2
`
`

`

`Regions within a protein chain or in separate chains sometimes are cross-linked covalently
`through disulfide bonds. Although disulfide bonds are rare in intracellular proteins, they are
`commonly found in extracellular proteins, where they help maintain the native, folded structure.
`The smallest amino acid, glycine, has a single hydrogen atom as its R group. Its small size allows
`it to fit into tight spaces. Unlike any of the other common amino acids, proline has a cyclic ring
`that is produced by formation of a covalent bond between its R group and the amino group on
`C . Proline is very rigid, and its presence creates a fixed kink in a protein chain. Proline and

`glycine are sometimes found at points on a protein’s surface where the chain loops back into the
`protein.
`
`The 6225 known and predicted proteins encoded by the yeast genome have an average molecular
`weight (MW) of 52,728 and contain, on average, 466 amino acid residues. Assuming that these
`average values represent a “typical” eukaryotic protein, then the average molecular weight of
`amino acids is 113, taking their average relative abundance in proteins into account. This is a
`useful number to remember, as we can use it to estimate the number of residues from the
`molecular weight of a protein or vice versa. Some amino acids are more abundant in proteins
`than other amino acids. Cysteine, tryptophan, and methionine are rare amino acids; together they
`constitute approximately 5 percent of the amino acids in a protein. Four amino acids—leucine,
`serine, lysine, and glutamic acid—are the most abundant amino acids, totaling 32 percent of all
`the amino acid residues in a typical protein. However, the amino acid composition of proteins
`can vary widely from these values. For example, as discussed in later sections, proteins that
`reside in the lipid bilayer are enriched in hydrophobic amino acids.
`
`Peptide Bonds Connect Amino Acids into Linear Chains
`Nature has evolved a single chemical linkage, the peptide bond, to connect amino acids into a
`linear, unbranched chain. The peptide bond is formed by a condensation reaction between the
`amino group of one amino acid and the carboxyl group of another (Figure 3-3a). The repeated
`amide N, C , and carbonyl C atoms of each amino acid residue form the backbone of a protein

`molecule from which the various side-chain groups project. As a consequence of the peptide
`linkage, the backbone has polarity, since all the amino groups lie to the same side of the C

`atoms. This leaves at opposite ends of the chain a free (unlinked) amino group (the N-terminus)
`and a free carboxyl group (the C-terminus). A protein chain is conventionally depicted with its
`N-terminal amino acid on the left and its C-terminal amino acid on the right (Figure 3-3b).
`
`Figure 3-3
`
`MYLAN INST. EXHIBIT 1091 PAGE 3
`
`MYLAN INST. EXHIBIT 1091 PAGE 3
`
`

`

`The peptide bond. (a) A condensation reaction between two amino acids forms
`the peptide bond, which links all the adjacent residues in a protein chain. (b)
`Side-chain groups (R) extend from the backbone of a protein chain, in which
`the amino N, α (more...)
`
`Many terms are used to denote the chains formed by polymerization of amino acids. A short
`chain of amino acids linked by peptide bonds and having a defined sequence is a peptide; longer
`peptides are referred to as polypeptides. Peptides generally contain fewer than 20–30 amino acid
`residues, whereas polypeptides contain as many as 4000 residues. We reserve the term protein
`for a polypeptide (or a complex of polypeptides) that has a threedimensional structure. It is
`implied that proteins and peptides represent natural products of a cell.
`
`The size of a protein or a polypeptide is reported as its mass in daltons (a dalton is 1 atomic mass
`unit) or as its molecular weight (a dimensionless number). For example, a 10,000-MW protein
`has a mass of 10,000 daltons (Da), or 10 kilodaltons (kDa). In the last section of this chapter, we
`will discuss different methods for measuring the sizes and other physical characteristics of
`proteins.
`
`Four Levels of Structure Determine the Shape of Proteins
`The structure of proteins commonly is described in terms of four hierarchical levels of
`organization. These levels are illustrated in Figure 3-4, which depicts the structure of
`hemagglutinin, a surface protein on the influenza virus. This protein binds to the surface of
`animal cells, including human cells, and is responsible for the infectivity of the flu virus.
`
`Figure 3-4
`
`Four levels of structure in hemagglutinin, which is a long
`multimeric molecule whose three identical subunits are each
`composed of two chains, HA and HA . (a) Primary structure
`1
`2
`is illustrated by the amino acid sequence of residues 68 –195
`(more...)
`
`The primary structure of a protein is the linear arrangement, or sequence, of amino acid residues
`that constitute the polypeptide chain.
`
`Secondary structure refers to the localized organization of parts of a polypeptide chain, which
`can assume several different spatial arrangements. A single polypeptide may exhibit all types of
`secondary structure. Without any stabilizing interactions, a polypeptide assumes a random-coil
`structure. However, when stabilizing hydrogen bonds form between certain residues, the
`backbone folds periodically into one of two geometric arrangements: an α helix, which is a
`spiral, rodlike structure, or a β sheet, a planar structure composed of alignments of two or more β
`strands, which are relatively short, fully extended segments of the backbone. Finally, U-shaped
`four-residue segments stabilized by hydrogen bonds between their arms are called turns. They
`are located at the surfaces of proteins and redirect the polypeptide chain toward the interior.
`(These structures will be discussed in greater detail later.)
`
`Tertiary structure, the next-higher level of structure, refers to the overall conformation of a
`
`MYLAN INST. EXHIBIT 1091 PAGE 4
`
`MYLAN INST. EXHIBIT 1091 PAGE 4
`
`

`

`polypeptide chain, that is, the three-dimensional arrangement of all the amino acids residues. In
`contrast to secondary structure, which is stabilized by hydrogen bonds, tertiary structure is
`stabilized by hydrophobic interactions between the nonpolar side chains and, in some proteins,
`by disulfide bonds. These stabilizing forces hold the α helices, β strands, turns, and random coils
`in a compact internal scaffold. Thus, a protein’s size and shape is dependent not only on its
`sequence but also on the number, size, and arrangement of its secondary structures. For proteins
`that consist of a single polypeptide chain, monomeric proteins, tertiary structure is the highest
`level of organization.
`
`Multimeric proteins contain two or more polypeptide chains, or subunits, held together by
`noncovalent bonds. Quaternary structure describes the number (stoichiometry) and relative
`positions of the subunits in a multimeric protein. Hemagglutinin is a trimer of three identical
`subunits; other multimeric proteins can be composed of any number of identical or different
`subunits.
`
`In a fashion similar to the hierarchy of structures that make up a protein, proteins themselves are
`part of a hierarchy of cellular structures. Proteins can associate into larger structures termed
`macromolecular assemblies. Examples of such macromolecular assemblies include the protein
`coat of a virus, a bundle of actin filaments, the nuclear pore complex, and other large
`submicroscopic objects. Macromolecular assemblies in turn combine with other cell biopolymers
`like lipids, carbohydrates, and nucleic acids to form complex cell organelles.
`
`Graphic Representations of Proteins Highlight Different Features
`Different ways of depicting proteins convey different types of information. The simplest way to
`represent three-dimensional structure is to trace the course of the backbone atoms with a solid
`line (Figure 3-5a); the most complex model shows the location of every atom (Figure 3-5b; see
`also Figure 2-1a). The former shows the overall organization of the polypeptide chain without
`consideration of the amino acid side chains; the latter details the interactions among atoms that
`form the backbone and that stabilize the protein’s conformation. Even though both views are
`useful, the elements of secondary structure are not easily discerned in them.
`
`Figure 3-5
`
`Various graphic representations of the structure of Ras, a
`guanine nucleotide–binding protein. Guanosine diphosphate,
`the substrate that is bound, is shown as a blue space-filling
`figure in parts (a)–(d). (a) The C trace of Ras, (more...)

`
`Another type of representation uses common shorthand symbols for depicting secondary
`structure, cylinders for α helices, arrows for β strands, and a flexible stringlike form for parts of
`the backbone without any regular structure (Figure 3-5c). This type of representation emphasizes
`the organization of the secondary structure of a protein, and various combinations of secondary
`structures are easily seen.
`
`However, none of these three ways of representing protein structure conveys much information
`about the protein surface, which is of interest because this is where other molecules bind to a
`protein. Computer analysis in which a water molecule is rolled around the surface of a protein
`can identify the atoms that are in contact with the watery environment. On this water-accessible
`
`MYLAN INST. EXHIBIT 1091 PAGE 5
`
`MYLAN INST. EXHIBIT 1091 PAGE 5
`
`

`

`surface, regions having a common chemical (hydrophobicity or hydrophilicity) and electrical
`(basic or acidic) character can be mapped. Such models show the texture of the protein surface
`and the distribution of charge, both of which are important parameters of binding sites (Figure 3-
`5d). This view represents a protein as seen by another molecule.
`
`Secondary Structures Are Crucial Elements of Protein Architecture
`In an average protein, 60 percent of the polypeptide chain exists as two regular secondary
`structures, α helices and β sheets; the remainder of the molecule is in random coils and turns.
`Thus, α helices and β sheets are the major internal supportive elements in proteins. In this
`section, we explore the forces that favor formation of secondary structures. In later sections, we
`examine how these structures can pack into larger arrays.
`
`The α Helix
`Polypeptide segments can assume a regular spiral, or helical, conformation, called the α helix. In
`this secondary structure, the carbonyl oxygen of each peptide bond is hydrogen-bonded to the
`amide hydrogen of the amino acid four residues toward the C-terminus. This uniform
`arrangement of bonds confers a polarity on a helix because all the hydrogen-bond donors have
`the same orientation. The peptide backbone twists into a helix having 3.6 amino acids per turn
`(Figure 3-6). The stable arrangement of amino acids in the α helix holds the backbone as a
`rodlike cylinder from which the side chains point outward. The hydrophobic or hydrophilic
`quality of the helix is determined entirely by the side chains, because the polar groups of the
`peptide backbone are already involved in hydrogen bonding in the helix and thus are unable to
`affect its hydrophobicity or hydrophilicity.
`
`Figure 3-6
`
`Model of the α helix. The polypeptide backbone is folded into
`a spiral that is held in place by hydrogen bonds (black dots)
`between backbone oxygen atoms and hydrogen atoms. Note
`that all the hydrogen bonds have the same polarity. The outer
`surface (more...)
`
`In many α helices hydrophilic side chains extend from one side of the helix and hydrophobic side
`chains from the opposite side, making the overall structure amphipathic. In such helices the
`hydrophobic residues, although apparently randomly arranged, occur in a regular pattern (Figure
`3-7). One way of visualizing this arrangement is to look down the center of an α helix and then
`project the amino acid residues onto the plane of the paper. The residues will appear as a wheel,
`and in the case of an amphipathic helix, the hydrophobic residues all lie on one side of the wheel
`and the hydrophilic ones on the other side.
`
`Figure 3-7
`
`Regions of an α helix may be amphipathic. The five chains of
`cartilage oligomeric matrix protein associate into a coiled-coil
`fibrous domain through amphipathic α helices. Seen in cross
`
`MYLAN INST. EXHIBIT 1091 PAGE 6
`
`MYLAN INST. EXHIBIT 1091 PAGE 6
`
`

`

`section through a part of the domain, the hy-drophobic
`(more...)
`
`Amphipathic α helices are important structural elements in fibrous proteins found in a watery
`environment. In a coiled-coil region of a protein, the hydrophobic surface of the α helix faces
`inward to form the hydrophobic core, and the hydrophilic surfaces face outward toward the
`surrounding fluid. This same orientation of surfaces is also found in most globular proteins. A
`crucial difference is that the hydrophobic interaction could be with a β strand, random coil, or
`another α helix. As we discuss later, amphipathic β strands line the walls of an ion channel in the
`cell membrane.
`
`The β Sheet
`Another regular secondary structure, the β sheet, consists of laterally packed β strands. Each β
`strand is a short (5–8-residue), nearly fully extended polypeptide chain. Hydrogen bonding
`between backbone atoms in adjacent β strands, within either the same or different polypeptide
`chains, forms a β sheet (Figure 3-8a). Like α helices, β strands have a polarity defined by the
`orientation of the peptide bond. Therefore, in a pleated sheet, adjacent β strands can be oriented
`antiparallel or parallel with respect to each other. In both arrangements of the backbone, the side
`chains project from both faces of the sheet (Figure 3-8b).
`
`Figure 3-8
`
`β SHEETS. (a) A simple two-stranded β sheet with
`antiparallel β strands. A sheet is stabilized by hydrogen bonds
`(black dots) between the β strands. The planarity of the
`peptide bond forces a β sheet to be pleated; (more...)
`
`In some proteins, β sheets form the floor of a binding pocket (Figure 3-8c). In many structural
`proteins, multiple layers of pleated sheets provide toughness. Silk fibers, for example, consist
`almost entirely of stacks of antiparallel β sheets. The fibers are flexible because the stacks of β
`sheets can slip over one another. However, they are also resistant to breakage because the peptide
`backbone is aligned parallel with the fiber axis.
`
`Turns
`Composed of three or four residues, turns are compact, U-shaped secondary structures stabilized
`by a hydrogen bond between their end residues. They are located on the surface of a protein,
`forming a sharp bend that redirects the polypeptide backbone back toward the interior. Glycine
`and proline are commonly present in turns. The lack of a large side chain in the case of glycine
`and the presence of a built-in bend in the case of proline allow the polypeptide backbone to fold
`into a tight U-shaped structure. Without turns, a protein would be large, extended, and loosely
`packed. A polypeptide backbone also may contain long bends, or loops. In contrast to turns,
`which exhibit a few defined structures, loops can be formed in many different ways.
`
`Motifs Are Regular Combinations of Secondary Structures
`Many proteins contain one or more motifs built from particular combinations of secondary
`
`MYLAN INST. EXHIBIT 1091 PAGE 7
`
`MYLAN INST. EXHIBIT 1091 PAGE 7
`
`

`

`structures. A motif is defined by a specific combination of secondary structures that has a
`particular topology and is organized into a characteristic three-dimensional structure. Three
`common motifs are depicted in Figure 3-9.
`
`Figure 3-9
`
`Secondary-structure motifs. (a) The coiled-coil motif (left) is
`characterized by two or more helices wound around one
`another. In some DNA-binding proteins, like c-Jun, a two-
`stranded coiled coil is responsible for dimerization (right).
`Each helix in (more...)
`
`The coiled-coil motif comprises two, three, or four amphipathic α helices wrapped around one
`another. In this motif, hydrophobic side chains project like “knobs” from one helix and
`interdigitate into the gaps, or “holes,” between the hydrophobic side chains of the other helix
`along the contact surface. The subunits in some multimeric proteins and in rodlike fibers are held
`2+
`together by coiled-coil interactions. The Ca
`-binding helix-loop-helix motif is marked by the
`presence of certain hydrophilic residues at invariant positions in the loop. Oxygen atoms in the
`invariant residues bind a calcium ion through hydrogen bonds. In another common motif, the
`zinc finger, three secondary structures—an α helix and two β strands with an antiparallel
`orientation—form a fingerlike bundle held together by a zinc ion. This motif is most commonly
`found in proteins that bind RNA or DNA.
`
`Additional motifs will be examined in discussions of other proteins. The presence of the same
`motif in different proteins with similar functions clearly indicates that during evolution these
`useful combinations of secondary structures have been conserved.
`
`Structural and Functional Domains Are Modules of Tertiary Structure
`The tertiary structure of large proteins is often subdivided into distinct globular or fibrous
`regions called domains. Structurally, a domain is a compactly folded region of polypeptide. For
`large proteins, domains can be recognized in structures determined by x-ray crystallography or in
`images captured by electron microscopy. These discrete regions are well distinguished or
`physically separated from other parts of the protein, but connected by the polypeptide chain.
`Hemagglutinin, for example, contains a globular domain and a fibrous domain (see Figure 3-4b).
`
`A structural domain consists of 100–200 residues in various combinations of α helices, β sheets,
`turns, and random coils. Often a domain is characterized by some interesting structural feature,
`for example, an unusual abundance of a particular amino acid (a proline-rich domain, an acidic
`domain, a glycine-rich domain), sequences common to (conserved in) many proteins (SH3, or
`Src homology region 3), or a particular secondary-structure motif (zinc-finger motif in kringle
`domain).
`
`Domains sometimes are defined in functional terms based on observations that the activity of a
`protein is localized to a small region along its length. For instance, a particular region or regions
`of a protein may be responsible for its catalytic activity (e.g., a kinase domain) or binding ability
`(e.g., a DNA-binding domain, membrane-binding domain). Functional domains often are
`identified experimentally by whittling down a protein to its smallest active fragment with the aid
`of proteases, enzymes that cleave the polypeptide backbone. Alternatively, the DNA encoding a
`
`MYLAN INST. EXHIBIT 1091 PAGE 8
`
`MYLAN INST. EXHIBIT 1091 PAGE 8
`
`

`

`protein can be subjected to mutagenesis, so that segments of the protein’s backbone are removed
`or changed (Chapter 7). The activity of the truncated or altered protein product synthesized from
`the mutated gene is then monitored.
`
`The functional definition of a domain is less rigorous than a structural definition. However, if the
`three-dimensional structure of a protein has not been determined, identification of functional
`domains can provide useful information about the protein. Because the activity of a protein
`usually depends on a proper three-dimensional structure, a functional domain consists of at least
`one and often several structural domains.
`
`The organization of tertiary structure into domains further illustrates the principle that complex
`molecules are built from simpler components. Like secondary-structure motifs, tertiary-structure
`domains are incorporated as modules into different proteins, thereby modifying their functional
`activities. The modular approach to protein architecture is particularly easy to recognize in large
`proteins, which tend to be a mosaic of different domains and thus can perform different functions
`simultaneously.
`
`The epidermal growth factor (EGF) domain is one example of a module that is present in several
`proteins (Figure 3-10). EGF is a small soluble peptide hormone that binds to cells in the skin and
`connective tissue, causing them to divide. It is generated by proteolytic cleavage between
`repeated EGF domains in the EGF precursor protein, which is anchored in the cell membrane by
`a membrane-spanning domain. Six conserved cysteine residues form three pairs of disulfide
`bonds that hold EGF in its native conformation. The EGF domain also occurs in other proteins,
`including tissue plasminogen activator (TPA), a protease that is used to dissolve blood clots in
`heart attack victims; Neu protein, which is involved in embryonic differentiation; and Notch
`protein, a cell-adhesion molecule that glues cells together. Besides the EGF domain, these
`proteins contain additional domains found in other proteins. For example, TPA possesses a
`chymotryptic domain, a common feature in proteins that catalyze proteolysis.
`
`Figure 3-10
`
`Schematic diagrams of various proteins, illustrating their
`modular nature. Epidermal growth factor (EGF) is generated
`by proteolytic cleavage of a precursor protein containing
`multiple EGF domains (orange). The EGF domain also
`occurs in Neu protein and (more...)
`
`Sequence Homology Suggests Functional and Evolutionary
`Relationships between Proteins
`Early evidence supporting the key principle that the amino acid sequence of a protein determines
`its three-dimensional structure was obtained in the 1960s by Max Perutz. On comparing the
`structures of myoglobin and hemoglobin determined from x-ray crystallographic analysis, he
`immediately noted that the subunits of hemoglobin, a tetramer of two α and two β subunits,
`resembled myoglobin, a monomer (Figure 3-11). Although the sequences of the two proteins
`were unknown at the time, Perutz proposed that the similar arrangement of α helices in the two
`proteins is a consequence of their having similar amino acid sequences. Later sequencing of
`myoglobin and hemoglobin revealed that many identical or chemically similar residues occur in
`identical positions throughout the sequences of both proteins. The two proteins also exhibit
`
`MYLAN INST. EXHIBIT 1091 PAGE 9
`
`MYLAN INST. EXHIBIT 1091 PAGE 9
`
`

`

`similar functions: myoglobin is the oxygen-carrier protein in muscle, and hemoglobin the
`oxygen-carrier protein in blood. Most of the conserved residues hold the heme group in place or
`are responsible for maintaining the hydrophobic interior of the protein.
`
`Figure 3-11
`
`Models of the tertiary structures of the oxygen-carrier
`proteins myoglobin and hemoglobin based on x-ray
`crystallographic analysis. Note the similarity in the tertiary
`structures of myoglobin and the two α subunits (blue) and
`two β subunits (more...)
`
`As data concerning protein sequences and three- dimensional structures accumulated, the
`concept that similar sequences fold into similar secondary and tertiary structures was confirmed.
`The propensity of each amino acid to occur in the various types of secondary structures has been
`calculated from the amino acid sequence of secondary structures extracted from databases of the
`three-dimensional structures of proteins. This tabulation of the folding information inherent in
`the sequence is now being used in attempts to predict the three-dimensional structure of various
`proteins from their amino acid sequences.
`
`In the classical taxonomy of the eighteenth and nineteenth centuries, organisms were classified
`according to their morphological similarities and differences. In this century, the molecular
`revolution in biology has given birth to “molecular” taxonomy: the classification of proteins
`based on similarities and differences in their amino acid sequences. This new taxonomy provides
`much information about protein function and evolutionary relationships. If the similarity between
`proteins from different organisms is significant over their entire sequence, then the proteins are
`homologs of one another, and they probably carry out similar functions. Sequence similarity also
`suggests an evolutionary relationship between proteins; that is, they evolved from a common
`ancestor. We can therefore describe homologous proteins as belonging to the same “family” and
`can trace their lineage from comparisons of sequences. Closely related proteins have the most
`similar sequences; distantly related proteins have only faintly similar sequences.
`
`The kinship among homologous proteins is most easily visualized from a tree diagram based on
`sequence analyses. For example, the amino acid sequences of hemoglobins from different
`species suggest that they evolved from an ancestral monomeric, oxygen-binding protein (Figure
`3-12). Over time, this ancestral protein slowly changed, giving rise to myoglobin, which
`remained a monomeric protein, and to the α and β subunits, which evolved to associate into the
`tetrameric hemoglobin molecule. As the tree diagram in Figure 3-12 shows, evolution of the
`globin protein family parallels that of the vertebrates.
`
`Figure 3-12
`
`Evolutionary tree showing how the globin protein family
`arose, starting from the most primitive oxygen-binding
`proteins, leghemoglobins, in plants. Sequence comparisons
`have revealed that evolution of the globin proteins parallels
`the evolution of vertebrates. (more...)
`
`MYLAN INST. EXHIBIT 1091 PAGE 10
`
`MYLAN INST. EXHIBIT 1091 PAGE 10
`
`

`

`The power of such comparative analysis and identification of homologous proteins has expanded
`substantially in recent years by use of the base sequences in an organism’s genome to deduce the
`amino acid sequences of the encoded proteins. As discussed in Chapter 7, this approach permits
`“sequencing” of proteins that are difficult to purify in significant amounts.
`
`SUMMARY
`
` A protein is a linear polymer of amino acids linked together by peptide bonds. Various,
`mostly noncovalent, interactions between amino acids in the linear sequence stabilize a
`specific folded three-dimensional structure (conformation) for each protein.
`
` The 20 different amino acids found in natural proteins are conveniently grouped into three
`categories based on the nature of their side (R) groups: hydrophilic amino acids, with a
`charged or polar and uncharged R group; hydrophobic amino acids, with an aliphatic or
`bulky and aromatic R group; and amino acids with a special group, consisting of cysteine,
`glycine, and proline (see Figure 3-2).
`
` The α helix, β strand and sheet, and turn are the most prevalent elements of protein
`secondary structure, which is stabilized by hydrogen bonds between atoms of the peptide
`backbone. Certain combinations of secondary structures give rise to different motifs,
`which are found in a variety of proteins and often are associated with specific functions
`(see Figure 3-9).
`
` Protein tertiary structure results from hydrophobic interactions and disulfide bonds that
`stabilize folding of the secondary structure into a compact overall arrangement, or
`conformation. Large proteins often contain distinct domains, independently folded

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