[go: up one dir, main page]

Next Issue
Volume 4, March
Previous Issue
Volume 3, September
 
 

Biomolecules, Volume 3, Issue 4 (December 2013) – 14 articles , Pages 733-1052

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
873 KiB  
Review
Protein Stability, Folding and Misfolding in Human PGK1 Deficiency
by Giovanna Valentini, Maristella Maggi and Angel L. Pey
Biomolecules 2013, 3(4), 1030-1052; https://doi.org/10.3390/biom3041030 - 18 Dec 2013
Cited by 16 | Viewed by 9320
Abstract
Conformational diseases are often caused by mutations, altering protein folding and stability in vivo. We review here our recent work on the effects of mutations on the human phosphoglycerate kinase 1 (hPGK1), with a particular focus on thermodynamics and kinetics of protein [...] Read more.
Conformational diseases are often caused by mutations, altering protein folding and stability in vivo. We review here our recent work on the effects of mutations on the human phosphoglycerate kinase 1 (hPGK1), with a particular focus on thermodynamics and kinetics of protein folding and misfolding. Expression analyses and in vitro biophysical studies indicate that disease-causing mutations enhance protein aggregation propensity. We found a strong correlation among protein aggregation propensity, thermodynamic stability, cooperativity and dynamics. Comparison of folding and unfolding properties with previous reports in PGKs from other species suggests that hPGK1 is very sensitive to mutations leading to enhance protein aggregation through changes in protein folding cooperativity and the structure of the relevant denaturation transition state for aggregation. Overall, we provide a mechanistic framework for protein misfolding of hPGK1, which is insightful to develop new therapeutic strategies aimed to target native state stability and foldability in hPGK1 deficient patients. Full article
(This article belongs to the Special Issue Protein Folding and Misfolding)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Superimposition of the C<sup>α</sup> traces of three PGK structures: <span class="html-italic">Escherichia coli</span> (blue, PDB ID 1ZMR), <span class="html-italic">Saccharomyces cerevisiae</span> (green, PDB ID 1FW8) and <span class="html-italic">Homo sapiens</span> (red, PDB ID 2XE7). The conformation of each individual domain, except for a few source-specific insertions, is similar in all structures. Both inter-domain helix and L hinge are almost completely superposed in all the three PGKs, pointing to a conservative evolution of the structural elements and underlining their essential role in the protein function. The bound ligands are shown as thin stick models.</p>
Full article ">Figure 2
<p>Ribbons representation of human PGK1 in open conformation (PDB ID, 2XE7). The bound ligands are shown as stick model. The side chains of amino acids affected by mutations are indicated by black spheres and represented as ball and stick.</p>
Full article ">Figure 3
<p>Aggregation of hPGK1 enzymes <span class="html-italic">in vitro</span> and upon expression in <span class="html-italic">E. coli</span> cultures (A) Fraction of soluble hPGK1 existing as a monomer upon expression analyses of hPGK1 enzymes in <span class="html-italic">E. coli</span>. The data are means ± s.d. from three independent expression experiments and were obtained starting from the same amount of total soluble protein and upon purification by ion-exchange and size exclusion chromatography as described [<a href="#B13-biomolecules-03-01030" class="html-bibr">13</a>,<a href="#B45-biomolecules-03-01030" class="html-bibr">45</a>] (<b>B</b>) Normalized thermal denaturation profiles for WT hPGK1 monitored by activity, light scattering (LS) and differential scanning calorimetry (DSC) measurements; Note that the three techniques provide similar denaturation profiles. (<b>C</b>) Arrhenius plots for the thermal denaturation of WT hPGK1 obtained from DSC experiments using a two-state irreversible model [<a href="#B13-biomolecules-03-01030" class="html-bibr">13</a>] and used to determine denaturation rates at 37 °C (indicated by the vertical dotted line).</p>
Full article ">Figure 4
<p>Plausible mechanisms linking the reduced kinetic stability (towards aggregation or proteolysis) and the lower unfolding cooperativity in disease-causing hPGK1 mutants. In the case of kinetically stable variants (panel <b>A</b>), the levels of the highly kinetically sensitive intermediate (I) levels are always very low, and aggregation and proteolysis occur mainly through the unfolded state U (left side of the figure). However, in some mutants (panel <b>B</b>), the population of the intermediate may raise (and can be “detected” by a low <span class="html-italic">m</span>-value in urea denaturations), leading to a significant contribution from the I state to the irreversible denaturation kinetics. It must be also noted that in panel B, at low denaturing stress, the I state is much more populated than the U state, and thus, may contribute much more to the kinetics of proteolysis and/or aggregation than the U state.</p>
Full article ">Figure 5
<p>A simple scenario for protein folding and misfolding hPGK1 enzymes inside the cell. Folding of hPGK1 may occur <span class="html-italic">in vivo</span> spontaneously after its synthesis, as seen by its spontaneous folding <span class="html-italic">in vitro</span> [<a href="#B13-biomolecules-03-01030" class="html-bibr">13</a>] and the fast and spontaneous folding of yeast PGK in living cells ([<a href="#B26-biomolecules-03-01030" class="html-bibr">26</a>,<a href="#B99-biomolecules-03-01030" class="html-bibr">99</a>]). Alternatively, the native state can be reached via interaction of partially folded states with different classes of molecular chaperones. The red arrows indicate steps that may be affected <span class="html-italic">in vivo</span> upon disease-causing mutations, mostly associated with reduced native stability and folding cooperativity [<a href="#B13-biomolecules-03-01030" class="html-bibr">13</a>,<a href="#B46-biomolecules-03-01030" class="html-bibr">46</a>]. The potential targets for pharmacological correction by proteostasis modulators (PM) and pharmacological chaperones (PC) are also indicated.</p>
Full article ">
501 KiB  
Review
Toxin Instability and Its Role in Toxin Translocation from the Endoplasmic Reticulum to the Cytosol
by Ken Teter
Biomolecules 2013, 3(4), 997-1029; https://doi.org/10.3390/biom3040997 - 10 Dec 2013
Cited by 18 | Viewed by 7504
Abstract
AB toxins enter a host cell by receptor-mediated endocytosis. The catalytic A chain then crosses the endosome or endoplasmic reticulum (ER) membrane to reach its cytosolic target. Dissociation of the A chain from the cell-binding B chain occurs before or during translocation to [...] Read more.
AB toxins enter a host cell by receptor-mediated endocytosis. The catalytic A chain then crosses the endosome or endoplasmic reticulum (ER) membrane to reach its cytosolic target. Dissociation of the A chain from the cell-binding B chain occurs before or during translocation to the cytosol, and only the A chain enters the cytosol. In some cases, AB subunit dissociation is facilitated by the unique physiology and function of the ER. The A chains of these ER-translocating toxins are stable within the architecture of the AB holotoxin, but toxin disassembly results in spontaneous or assisted unfolding of the isolated A chain. This unfolding event places the A chain in a translocation-competent conformation that promotes its export to the cytosol through the quality control mechanism of ER-associated degradation. A lack of lysine residues for ubiquitin conjugation protects the exported A chain from degradation by the ubiquitin-proteasome system, and an interaction with host factors allows the cytosolic toxin to regain a folded, active state. The intrinsic instability of the toxin A chain thus influences multiple steps of the intoxication process. This review will focus on the host–toxin interactions involved with A chain unfolding in the ER and A chain refolding in the cytosol. Full article
(This article belongs to the Special Issue Protein Folding and Misfolding)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Intracellular toxin trafficking. The general trafficking and translocation itinerary for AB-type, endoplasmic reticulum (ER)-translocating toxins is shown. These toxins bind to distinct surface receptors and are internalized by a variety of endocytic mechanisms. The internalized toxin is recycled to the plasma membrane, directed to the lysosomes for degradation, or delivered to the trans-Golgi network (TGN) en route to the ER translocation site. Vesicle-mediated transport to the TGN can originate from the early or late endosomes, depending on which toxin is present. Likewise, multiple retrograde transport pathways can deliver the toxin from the TGN to the ER. The toxin may cycle between the Golgi and ER until the catalytic subunit dissociates from the rest of the toxin and shifts to an unfolded conformation which triggers its export to the cytosol in a process involving the quality control system of ER-associated degradation. Some of the free, ER-localized A chain escapes ER-associated degradation (ERAD) and is secreted back into the medium via Golgi and TGN intermediates. In most cell types, trafficking from the cell surface to the ER is very inefficient: the majority of internalized toxin is routed to the lysosomes, and only around 10% of surface-bound toxin reaches the ER [<a href="#B9-biomolecules-03-00997" class="html-bibr">9</a>,<a href="#B10-biomolecules-03-00997" class="html-bibr">10</a>,<a href="#B11-biomolecules-03-00997" class="html-bibr">11</a>,<a href="#B12-biomolecules-03-00997" class="html-bibr">12</a>,<a href="#B13-biomolecules-03-00997" class="html-bibr">13</a>,<a href="#B14-biomolecules-03-00997" class="html-bibr">14</a>,<a href="#B15-biomolecules-03-00997" class="html-bibr">15</a>,<a href="#B16-biomolecules-03-00997" class="html-bibr">16</a>]. Thus, ectopic expression of an ER-localized A chain via transfected cultured cells, transformed yeast, or microsomal transcription/translation systems is often used for toxin translocation studies.</p>
Full article ">Figure 2
<p>Structural organization of AB-type, ER-translocating toxins. (<b>a</b>) Ribbon diagram of cholera toxin (Ctx; PDB 1S5F, [<a href="#B33-biomolecules-03-00997" class="html-bibr">33</a>]). The A1 subunit is in blue; the A2 linker is red; and the B homopentamer is grey. The CtxB pentamer recognizes GM1 gangliosides on the host cell surface, while CtxA1 is an ADP-ribosyltransferase that elevates intracellular cAMP levels by activating the stimulatory α subunit of the heterotrimeric G protein; (<b>b</b>) Ribbon diagram of ricin toxin (Rtx; PDB 2AAI, [<a href="#B34-biomolecules-03-00997" class="html-bibr">34</a>]). RtxA is in blue, and RtxB is in grey. RtxB binds to a wide range of glycoproteins and glycolipids with terminal galactose residues, while RtxA is an <span class="html-italic">N</span>-glycosidase that inhibits protein synthesis by removing a specific adenine residue from the 28S rRNA; (<b>c</b>) Ribbon diagram of pertussis toxin (Ptx; PDB 1PRT, [<a href="#B35-biomolecules-03-00997" class="html-bibr">35</a>]). The catalytic S1 subunit is in blue, and the five subunits of the B pentamer (S2, S3, two copies of S4, and S5) are grey. PtxB can bind to a variety of glycoconjugates, while PtxS1 is an ADP-ribosyltransferase that elevates intracellular cAMP levels by locking the inhibitory α subunit of the heterotrimeric G protein in an inactive state; (<b>d</b>) Ribbon diagram of Shiga toxin (Stx; PDB 1DM0, [<a href="#B36-biomolecules-03-00997" class="html-bibr">36</a>]). The A1 subunit is in blue; the A2 linker is in red; and the B homopentamer is grey. The StxB pentamer binds to globoside Gb3 on the host cell surface, while StxA1 is an <span class="html-italic">N</span>-glycosidase that inhibits protein synthesis by removing a specific adenine residue from the 28S rRNA. The Stx family includes Stx from <span class="html-italic">Shigella dysenteriae</span> (pictured) and the Shiga-like toxins (Stx1, Stx2, and Stx2 isoforms) from <span class="html-italic">Escherichia coli</span>; (<b>e</b>) Ribbon diagram of <span class="html-italic">Pseudomonas aeruginosa</span> exotoxin A (EtxA; PDB 1IKQ, [<a href="#B37-biomolecules-03-00997" class="html-bibr">37</a>]). The catalytic moiety (domain III) is in blue, and the B moiety (domains I and II) is in grey. The B moiety of EtxA binds to the α-macroglobulin receptor/low density lipoprotein receptor-related protein on the host plasma membrane, while the A moiety of EtxA is an ADP-ribosyltransferase that inhibits protein synthesis through the modification of elongation factor 2; (<b>f</b>) Ribbon diagram of cytolethal distending toxin (Cdtx; PDB 1SR4, [<a href="#B38-biomolecules-03-00997" class="html-bibr">38</a>]). The catalytic CdtxB subunit is in blue, while the cell-binding CdtxA and CdtxC subunits are in grey. The cell-binding heterodimer binds to cholesterol and glycoconjugates, while the CdtxB subunit is a type I DNase that induces cell cycle arrest by causing double-stranded DNA breaks.</p>
Full article ">
1336 KiB  
Article
Control of Collagen Stability and Heterotrimer Specificity through Repulsive Electrostatic Interactions
by Avanish S. Parmar, Mihir Joshi, Patrick L. Nosker, Nida F. Hasan and Vikas Nanda
Biomolecules 2013, 3(4), 986-996; https://doi.org/10.3390/biom3040986 - 6 Dec 2013
Cited by 8 | Viewed by 6765
Abstract
Charge-pair interactions between acidic and basic residues on the surface of collagen can promote stability as well as control specificity of molecular recognition. Heterotrimeric collagen peptides have been engineered de novo using either rational or computational methods, which in both cases optimize networks [...] Read more.
Charge-pair interactions between acidic and basic residues on the surface of collagen can promote stability as well as control specificity of molecular recognition. Heterotrimeric collagen peptides have been engineered de novo using either rational or computational methods, which in both cases optimize networks of favorable charge-pair interactions in the target structure. Less understood is the role of electrostatic repulsion between groups of like charge in destabilizing structure or directing molecular recognition. To study this, we apply a “charge crowding” approach, where repulsive interactions between multiple aspartate side chains are found to destabilize the homotrimer states in triple helical peptide system and can be utilized to promote the formation of heterotrimers. Neutralizing surface charge by increasing salt concentration or decreasing pH can enhance homotrimer stability, confirming the role of charge crowding on the destabilization of homotrimers via electrostatic repulsion. Charge crowding may be used in conjunction with other approaches to create specific collagen heterotrimers. Full article
(This article belongs to the Special Issue Protein Folding and Misfolding)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Circular Dichroism (CD) spectra at 4 °C of all 10 mixtures of peptides A, B, C.</p>
Full article ">Figure 2
<p>Thermal denaturation (top row) and their respective first derivative plots (bottom row) for peptide mixtures (<b>A</b> and <b>D</b>) 3A, 3B, 3C (<b>B</b> and <b>E</b>) A:B:C, 2A:C, A:2C (<b>C</b> and <b>F</b>) 2A:B, A:2B, 2B:C, B:2C.</p>
Full article ">Figure 3
<p>NMR measurement of <sup>1</sup>H-<sup>13</sup>C HSQC spectra. (<b>A</b>) Merged spectra of A and B alone (black) <span class="html-italic">versus</span> A:B mixture (red); (<b>B</b>) Merged spectra of A and C alone (black) <span class="html-italic">versus</span> A:C mixture (red); (<b>C</b>) Merged spectra of B and C alone (black) <span class="html-italic">versus</span> B:C mixture (red); (<b>D</b>) Merged spectra of A, B, C, A:C (all black) <span class="html-italic">versus</span> A:B:C mixture (red).</p>
Full article ">Figure 4
<p>Thermal denaturation (top row) and their respective first derivative plots (bottom row) for A, B and C at various salt concentrations.</p>
Full article ">Figure 5
<p>Thermal denaturation (top row) and their respective first derivative plots (bottom row) for A, B and C pH7 and pH2.5.</p>
Full article ">Figure 6
<p>Sequence of fibrous natural collagens. Glycine is marked in yellow, clusters of like charges in red, clusters of favorable interactions in blue.</p>
Full article ">
961 KiB  
Article
Unfolding Thermodynamics of Cysteine-Rich Proteins and Molecular Thermal-Adaptation of Marine Ciliates
by Giorgia Cazzolli, Tatjana Škrbić, Graziano Guella and Pietro Faccioli
Biomolecules 2013, 3(4), 967-985; https://doi.org/10.3390/biom3040967 - 18 Nov 2013
Cited by 5 | Viewed by 6377
Abstract
Euplotes nobilii and Euplotes raikovi are phylogenetically closely allied species of marine ciliates, living in polar and temperate waters, respectively. Their evolutional relation and the sharply different temperatures of their natural environments make them ideal organisms to investigate thermal-adaptation. We perform a comparative [...] Read more.
Euplotes nobilii and Euplotes raikovi are phylogenetically closely allied species of marine ciliates, living in polar and temperate waters, respectively. Their evolutional relation and the sharply different temperatures of their natural environments make them ideal organisms to investigate thermal-adaptation. We perform a comparative study of the thermal unfolding of disulfide-rich protein pheromones produced by these ciliates. Recent circular dichroism (CD) measurements have shown that the two psychrophilic (E. nobilii) and mesophilic (E. raikovi) protein families are characterized by very different melting temperatures, despite their close structural homology. The enhanced thermal stability of the E. raikovi pheromones is realized notwithstanding the fact that these proteins form, as a rule, a smaller number of disulfide bonds. We perform Monte Carlo (MC) simulations in a structure-based coarse-grained (CG) model to show that the higher stability of the E. raikovi pheromones is due to the lower locality of the disulfide bonds, which yields a lower entropy increase in the unfolding process. Our study suggests that the higher stability of the mesophilic E. raikovi phermones is not mainly due to the presence of a strongly hydrophobic core, as it was proposed in the literature. In addition, we argue that the molecular adaptation of these ciliates may have occurred from cold to warm, and not from warm to cold. To provide a testable prediction, we identify a point-mutation of an E. nobilii pheromone that should lead to an unfolding temperature typical of that of E. raikovi pheromones. Full article
(This article belongs to the Special Issue Protein Folding and Misfolding)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Ribbon presentations of the three-dimensional native structure of the E<span class="html-italic">r</span>-1 (<b>A</b>) and E<span class="html-italic">n</span>-1 (<b>B</b>) pheromones.</p>
Full article ">Figure 2
<p>Comparison of the Cys-Cys bond patterns in the native E<span class="html-italic">r</span>-1 (<b>A</b>) and E<span class="html-italic">n</span>-1 (<b>B</b>) pheromone structures. Spheres of identical color indicate cysteine residues paired together into a disulfide bond.</p>
Full article ">Figure 3
<p>Far-UV circular dichroism (CD) spectra (expressed in <span class="html-italic">Δε</span> units) of the mesophilic, E<span class="html-italic">r</span>-1 (<b>A</b>), and psychrophilic, E<span class="html-italic">n</span>-1 (<b>B</b>), pheromones (20 mM, pH 6) at three selected temperatures (red 40 °C; green 70 °C; and blue 90 °C). The raw CD data were kindly supplied by M. Geralt and reported in [<a href="#B2-biomolecules-03-00967" class="html-bibr">2</a>].</p>
Full article ">Figure 4
<p>Comparison of the temperature-dependence of the fluctuation of fraction of native contacts defined in Equation (1) for different E<span class="html-italic">r</span> (<b>A</b>) and E<span class="html-italic">n</span> (<b>B</b>) pheromones obtained by means of MC simulations in the CG model defined in the Experimental Section.</p>
Full article ">Figure 5
<p>Temperature dependence of the fraction of native contacts for different E<span class="html-italic">r</span> (<b>A</b>) and E<span class="html-italic">n</span> (<b>B</b>) pheromones, obtained by means of MC simulations in the CG model defined in the Experimental Section.</p>
Full article ">Figure 6
<p>Equilibrium distributions of the fraction of native contacts at different temperatures for the E<span class="html-italic">r</span>-1 (<b>A</b>) and E<span class="html-italic">n</span>-1 (<b>B</b>) pheromones.</p>
Full article ">Figure 7
<p>Temperature dependence of the fluctuation in the number of native contacts defined in Equation (1), for the E<span class="html-italic">n</span> pheromones and the mutant of the E<span class="html-italic">n</span>-1 chain. In the mutant, the characteristic peak signaling the unfolding transition disappears.</p>
Full article ">Figure 8
<p>Schematic representations of the two homologous proteins considered in the analytically solvable toy-model.</p>
Full article ">
1075 KiB  
Review
Control of Cell Differentiation by Mitochondria, Typically Evidenced in Dictyostelium Development
by Yasuo Maeda and Junji Chida
Biomolecules 2013, 3(4), 943-966; https://doi.org/10.3390/biom3040943 - 11 Nov 2013
Cited by 19 | Viewed by 11036
Abstract
In eukaryotic cells, mitochondria are self-reproducing organelles with their own DNA and they play a central role in adenosine triphosphate (ATP) synthesis by respiration. Increasing evidence indicates that mitochondria also have critical and multiple functions in the initiation of cell differentiation, cell-type determination, [...] Read more.
In eukaryotic cells, mitochondria are self-reproducing organelles with their own DNA and they play a central role in adenosine triphosphate (ATP) synthesis by respiration. Increasing evidence indicates that mitochondria also have critical and multiple functions in the initiation of cell differentiation, cell-type determination, cell movement, and pattern formation. This has been most strikingly realized in development of the cellular slime mold Dictyostelium. For example, the expression of the mitochondrial ribosomal protein S4 (mt-rps4) gene is required for the initial differentiation. The Dictyostelium homologue (Dd-TRAP1) of TRAP-1 (tumor necrosis receptor-associated protein 1), a mitochondrial molecular chaperone belonging to the Hsp90 family, allows the prompt transition of cells from growth to differentiation through a novel prestarvation factor (PSF-3) in growth medium. Moreover, a cell-type-specific organelle named a prespore-specific vacuole (PSV) is constructed by mitochondrial transformation with the help of the Golgi complex. Mitochondria are also closely involved in a variety of cellular activities including CN-resistant respiration and apoptosis. These mitochondrial functions are reviewed in this article, with special emphasis on the regulation of Dictyostelium development. Full article
(This article belongs to the Special Issue Focus Update in Biomolecules)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The life cycle of <span class="html-italic">Dictyostelium discoideum</span> axenic strain Ax-2. The vegetative cells are usually grown in liquid medium, by means of pinocytotic incorporation of external nutrients. Under natural conditions, its parental strain <span class="html-italic">D. discoideum</span> NC-4 grows and multiplies by mitosis at the vegetative phase, phagocytosing nearby bacteria such as <span class="html-italic">Escherichia coli</span> and <span class="html-italic">Klebsiella aerogenes</span>. Upon exhaustion of nutrients, however, starving cells initiate differentiation, form multicellular structures (aggregates; mounds), and undergo a series of well-organized morphogenesis to construct fruiting bodies, each of which is consisting of a mass of spores (sorus) and a supporting cellular stalk. (see the text for details).</p>
Full article ">Figure 2
<p>A growth/differentiation checkpoint (GDT point) in the cell cycle of a <span class="html-italic">Dictyostelium discoideum</span> Ax-2 cell. The doubling time of axenically growing Ax-2 cells is about 7.2 h and most of their cell cycle is composed of G2-phase with little or no G1-phase and a short period of M- and S-phases. A specific checkpoint (referred to as the GDT point) of GDT is located at the mid–late G2-phase (just after T7 and just before T0). Ax-2 cells progress through their cell cycle to the GDT point, irrespective of the presence or absence of nutrients, and enter the differentiation phase from this point under starvation conditions [<a href="#B2-biomolecules-03-00943" class="html-bibr">2</a>]. T0, T1, and T7 indicates 0, 1, and 7 h, respectively, after a temperature shift from 11.5 °C to 22.0 °C for cell synchrony. The absence of G1 phase in the <span class="html-italic">Dictyostelium</span> cell cycle is not so strange, because there is little or no G1 phase in rapidly dividing cells such as animal cells at the cleavage stage, and also in the true slime mold <span class="html-italic">Physalum</span> and <span class="html-italic">Hydra</span>. (Basically from Maeda [<a href="#B3-biomolecules-03-00943" class="html-bibr">3</a>]).</p>
Full article ">Figure 3
<p>Strategy for creating <span class="html-italic">rps4</span>-null cells and their phenotypes. (<b>a</b>) As a starting material, LpCSfo cells in which pCoxIV (MTS)-<span class="html-italic">Sfo</span>I is expressed under the tetracycline-minus (−Tet) condition, were prepared. Subsequently, the mutant <span class="html-italic">rps4</span> gene (Mut-mtDNA for homologous recombination), in which the upstream <span class="html-italic">Sfo</span>I-site and a 5'-half of <span class="html-italic">rps4</span> coding region were deleted, was introduced into LpCSfo cells to obtain heteroplasmic transformants (LpCSfo<sup>HR</sup> cells) with mitochondria consisting of the Mut-mtDNA and wild-type mtDNA (Wt-mtDNA). Coupled with removal of Tet from growth medium, the fusion protein MTS-<span class="html-italic">Sfo</span>I synthesized in the cytoplasm is exclusively transferred into mitochondria of LpCSfo<sup>HR</sup> cells and selectively digests Wt-mtDNA but not Mut-mtDNA. Since the digested Wt-mtDNA is not duplicated, the Mut-mtDNA becomes dominant during the course of growth under the −Tet condition, thus eventually giving <span class="html-italic">rps4</span>-null cells. (<b>b</b>) These cells were grown in growth medium with (+Tet) or without (−Tet) teteracyclin. Membrane potential of mitochondria was visualized by staining of cells with MitoTracker Orange. As was expected, the staining of mitochondria was almost completely vanished in LpCSfo cells grown without Tet, because their mtDNA with an intact <span class="html-italic">Sfo</span>I site would be cleaved by <span class="html-italic">Sfo</span>I eventually to become a ρ<sup>0</sup> state devoid of mitochondrial DNA. Bars, 200 nm. (<b>c</b>) Development of starved MB35 cells and LpCSfo<sup>HR</sup> cells on agar. MB35 cells and LpCSfo<sup>HR</sup> cells grown with (+Tet) or without (−Tet) tetracycline were washed twice in BSS and plated on 1.5% non-nutrient agar at a density of 5 × 10<sup>6</sup> cells/cm<sup>2</sup>. This was followed by incubation for the indicated time at 22 °C. Bars, 0.5 mm. (Basically from Chida <span class="html-italic">et al</span>. [<a href="#B22-biomolecules-03-00943" class="html-bibr">22</a>]).</p>
Full article ">Figure 4
<p>Schematic drawing showing the behavior of Dd-TRAP1 during the prestarvation response (PSR) and the initiation of differentiation in <span class="html-italic">Dictyostelium</span> cells (See the text for details).</p>
Full article ">Figure 5
<p>A working hypothesis for explaining the existence of the GDT-point in the <span class="html-italic">Dictyostelium</span> cell-cycle, by assuming a temporally oscillating PSR activity during synchronized cell growth. This model is based on the fact that there is a good correlation between the expression patterns of the prestarvation (PS) genes and the GDT point-specific genes, and is also constructed from the presumption that the amount of PSFs (<span class="html-italic">i.e</span>., PSF activity) in an ideally synchronized cell population may oscillate in a cell-cycle dependent manner during growth, as shown here. The cell-cycle positions of T0 and T7 offer themselves repeatedly during the course of a completely synchronized growth (see <a href="#biomolecules-03-00943-f002" class="html-fig">Figure 2</a>). The PSR activity reaches the maximum value at the GDT-point (arrow between T7 and T0) in each cell cycle. Although it is presently unknown whether the expressions of Dd-TRAP1 and/or the amount of PSF-3 actually oscillate in a cell-cycle dependent manner, increasing their basal levels during the progression of synchronous cell growth, this issue remains to be examined in future. The horizontal axis of this figure represents incubation time (h) of an ideally synchronized <span class="html-italic">D. discoideum</span> (Ax-2) cell population in growth medium at 22.0 °C. T0 and T7 indicates 0 and 7 h, respectively, after a temperature shift from 11.5 °C to 22.0 °C for cell synchrony, and these are repeatedly represented during a prolonged time of synchronized culture.</p>
Full article ">Figure 6
<p>Electron-micrographs of PSV-mitochondrion complexes present in a 47.5%–50% fraction (the intermediate fraction between a lighter pure mitochondria and a heavier pure PSV fractions) of a multilayered sucrose gradient for cell fractionation. (<b>A</b>,<b>B</b>) the outer membrane of a mitochondrion (M) is continuous with the unit membrane of a PSV. (<b>C</b>) a PSV and a mitochondrion are partitioned by a single membrane derived from either the PSV or the mitochondrion, and the electron-dense lining membrane of the PSV is not observed at the contact region. (<b>D</b>) a PSV appears twisted at the contact region (arrow). (<b>E</b>) a completely formed PSV is in close contact with a mitochondrion. (<b>F</b>) a mitochondrial part itself loses its inner structural integrity and seems to be transformed into the PSV. Bar, 500 nm. (From Maeda [<a href="#B49-biomolecules-03-00943" class="html-bibr">49</a>]).</p>
Full article ">Figure 7
<p>A diagrammatic representation showing formation of the prespore-specific vacuole (PSV) from a mitochondrion and Golgi vesicles. Prior to PSV formation, mitochondria in differentiating prespore cells undergo drastic transformation to form a sort of vacuole (M vacuole), and some Dd-TRAP1 molecules translocate into the M vacuole. Subsequently, Golgi vesicles containing DIA2, Dd-GRP94 and other materials required for PSV formation fuse with the M vacuole, thus resulting in formation of the lining membrane (red) and the internal fibrous structure in the M vacuole. The mitochondrion-PSV complex is eventually twisted at the junction (arrow) and detached to form the respective organelles. Interestingly, almost all of the DIA2 molecules are selectively translocated to PSVs (possibly M vacuoles) in differentiating prespore cells, and seem to be required for exocytotic secretion of PSVs to form the outer-most membrane of spore cell wall [<a href="#B15-biomolecules-03-00943" class="html-bibr">15</a>]. M-V, M vacuole; GRP94, glucose-regulated protein 94 (endoplasmic reticulum Hsp90). (Slightly modified from Maeda [<a href="#B12-biomolecules-03-00943" class="html-bibr">12</a>]).</p>
Full article ">
659 KiB  
Review
Research Applications of Proteolytic Enzymes in Molecular Biology
by János András Mótyán, Ferenc Tóth and József Tőzsér
Biomolecules 2013, 3(4), 923-942; https://doi.org/10.3390/biom3040923 - 8 Nov 2013
Cited by 144 | Viewed by 27883
Abstract
Proteolytic enzymes (also termed peptidases, proteases and proteinases) are capable of hydrolyzing peptide bonds in proteins. They can be found in all living organisms, from viruses to animals and humans. Proteolytic enzymes have great medical and pharmaceutical importance due to their key role [...] Read more.
Proteolytic enzymes (also termed peptidases, proteases and proteinases) are capable of hydrolyzing peptide bonds in proteins. They can be found in all living organisms, from viruses to animals and humans. Proteolytic enzymes have great medical and pharmaceutical importance due to their key role in biological processes and in the life-cycle of many pathogens. Proteases are extensively applied enzymes in several sectors of industry and biotechnology, furthermore, numerous research applications require their use, including production of Klenow fragments, peptide synthesis, digestion of unwanted proteins during nucleic acid purification, cell culturing and tissue dissociation, preparation of recombinant antibody fragments for research, diagnostics and therapy, exploration of the structure-function relationships by structural studies, removal of affinity tags from fusion proteins in recombinant protein techniques, peptide sequencing and proteolytic digestion of proteins in proteomics. The aim of this paper is to review the molecular biological aspects of proteolytic enzymes and summarize their applications in the life sciences. Full article
(This article belongs to the Special Issue Enzymes and Their Biotechnological Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Action of aminopeptidases and carboxypeptidases removing the terminal amino acid residues as well as endopeptidases on a polypeptide substrate (having n residues). Red arrows show the peptide bonds to be cleaved.</p>
Full article ">Figure 2
<p>Kinetically-controlled synthesis of Z-<span class="html-small-caps">d</span>-Leu-<span class="html-small-caps">l</span>-Leu-NH<sub>2</sub> dipeptide. After the formation of enzyme-substrate complex (K<sub>1</sub>) a covalent enzyme-substrate intermediate is formed (K<sub>2</sub>). The intermediate is subjected to the attack from H<sub>2</sub>O or other nucleophiles (Nu). K<sub>H</sub> is the equilibrium constant of hydrolysis, K<sub>T</sub> is the equilibrium constant of the transferase reaction.</p>
Full article ">Figure 3
<p>Structure of IgG antibody molecules (<b>A</b>) and fragments released after proteolytic digestion using papain (<b>B</b>), pepsin (<b>C</b>) or ficin (<b>D</b>).</p>
Full article ">Figure 4
<p>Steps of proteomic analysis using mass-spectrometry after separation and in-gel digestion of proteins of interest.</p>
Full article ">
778 KiB  
Article
Structural Evidence for the Tetrameric Assembly of Chemokine CCL11 and the Glycosaminoglycan Arixtra™
by Andrew B. Dykstra, Matt D. Sweeney and Julie A. Leary
Biomolecules 2013, 3(4), 905-922; https://doi.org/10.3390/biom3040905 - 6 Nov 2013
Cited by 4 | Viewed by 7576
Abstract
Understanding chemokine interactions with glycosaminoglycans (GAG) is critical as these interactions have been linked to a number of inflammatory medical conditions, such as arthritis and asthma. To better characterize in vivo protein function, comprehensive knowledge of multimeric species, formed by chemokines under native [...] Read more.
Understanding chemokine interactions with glycosaminoglycans (GAG) is critical as these interactions have been linked to a number of inflammatory medical conditions, such as arthritis and asthma. To better characterize in vivo protein function, comprehensive knowledge of multimeric species, formed by chemokines under native conditions, is necessary. Herein is the first report of a tetrameric assembly of the human chemokine CCL11, which was shown bound to the GAG Arixtra™. Isothermal titration calorimetry data indicated that CCL11 interacts with Arixtra, and ion mobility mass spectrometry (IM-MS) was used to identify ions corresponding to the CCL11 tetrameric species bound to Arixtra. Collisional cross sections (CCS) of the CCL11 tetramer-Arixtra noncovalent complex were compared to theoretical CCS values calculated using a preliminary structure of the complex deduced using X-ray crystallography. Experimental CCS values were in agreement with theoretical values, strengthening the IM-MS evidence for the formation of the noncovalent complex. Tandem mass spectrometry data of the complex indicated that the tetramer-GAG complex dissociates into a monomer and a trimer-GAG species, suggesting that two CC-like dimers are bridged by Arixtra. As development of chemokine inhibitors is of utmost importance to treatment of medical inflammatory conditions, these results provide vital insights into chemokine-GAG interactions. Full article
(This article belongs to the Special Issue Focus Update in Biomolecules)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mass spectra of CCL11 under (<b>A</b>) denaturing conditions (50% acetonitrile, 0.1% formic acid) and (<b>B</b>) native conditions (250 mM ammonium acetate). M = monomer; D = dimer; Tr = Trimer; Te = Tetramer.</p>
Full article ">Figure 2
<p>(<b>A</b>) Mass spectra of (<b>A</b>) CCL11 and (<b>B</b>) CCL11+Arixtra magnifying the 3,000–4,500 <span class="html-italic">m/z</span> range by a factor of 93 (<b>A</b>) and 145 (<b>B</b>), respectively. Formation of the CCL11 tetramer-Arixtra complex is supported by the presence of peaks corresponding to multiple charge states of the complex at 3,178, 3,584, 3,885, and 4,370 <span class="html-italic">m/z</span>. Higher order oligomers were not observed. D = dimer; Tr = trimer; Te = tetramer; * = species bound to Arixtra.</p>
Full article ">Figure 3
<p>Heat map plotting <span class="html-italic">m/z</span> as a function of drift time for the CCL11+Arixtra mixture. Mass range is limited to the 3,000–4,500 <span class="html-italic">m/z</span> range illustrated in <a href="#biomolecules-03-00905-f002" class="html-fig">Figure 2</a>. Ion mobility separation allowed for the characterization of the peak at <span class="html-italic">m/z</span> 3,345 as both [D]<sup>5+</sup> and [Te]<sup>10+</sup> species. D = dimer; Tr = trimer; Te = tetramer; * = species bound to Arixtra.</p>
Full article ">Figure 4
<p>Tandem mass spectrometry of the [Te*]<sup>12+</sup> peak at 2,913 <span class="html-italic">m/z</span> showing that as the Te* complex is fragmented, it dissociates into Tr* and M species. M = monomer; D = dimer; Tr = trimer; * = species bound to Arixtra.</p>
Full article ">Figure 5
<p>ATDs corresponding to the 8+–11+ charge states of the CCL11 tetramer+Arixtra complex.</p>
Full article ">Figure 6
<p>Preliminary X-ray crystallography structure of the CCL11 tetramer-Arixtra complex (shown in color) overlaying the structure of the CCL2 tetramer (1DOL, shown in gray). The Arixtra molecule is shown top center. CCL11 subunits one and two are lower right and upper right, respectively, while subunits three and four are lower left and upper left, respectively. CXC-type dimers are vertically related while CC-type dimers are diagonally related.</p>
Full article ">
425 KiB  
Review
Biophysical Insights into the Inhibitory Mechanism of Non-Nucleoside HIV-1 Reverse Transcriptase Inhibitors
by Grant Schauer, Sanford Leuba and Nicolas Sluis-Cremer
Biomolecules 2013, 3(4), 889-904; https://doi.org/10.3390/biom3040889 - 1 Nov 2013
Cited by 5 | Viewed by 10029
Abstract
HIV-1 reverse transcriptase (RT) plays a central role in HIV infection. Current United States Federal Drug Administration (USFDA)-approved antiretroviral therapies can include one of five approved non-nucleoside RT inhibitors (NNRTIs), which are potent inhibitors of RT activity. Despite their crucial clinical role in [...] Read more.
HIV-1 reverse transcriptase (RT) plays a central role in HIV infection. Current United States Federal Drug Administration (USFDA)-approved antiretroviral therapies can include one of five approved non-nucleoside RT inhibitors (NNRTIs), which are potent inhibitors of RT activity. Despite their crucial clinical role in treating and preventing HIV-1 infection, their mechanism of action remains elusive. In this review, we introduce RT and highlight major advances from experimental and computational biophysical experiments toward an understanding of RT function and the inhibitory mechanism(s) of NNRTIs. Full article
(This article belongs to the Special Issue Enzymes and Their Biotechnological Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Reverse transcription. Schematic of the multistep process of the conversion of viral RNA (red) into integrase-competent dsDNA (bottom) for insertion into the human genome. PBS, primer binding site.</p>
Full article ">Figure 2
<p>Structure of reverse transcriptase (RT). RT is pictured as a solvent-accessible surface area model (taken from PDB ID: 1RTD). The DNA/DNA template/primer is shown as a cartoon. The p51 subdomain is colored in grey, and the p66 subdomain is subdivided into thumb (green), fingers (red), palm (purple), connection (wheat) and RNase H (blue) domains. The polymerase and RNase H active site residues are colored in yellow.</p>
Full article ">Figure 3
<p>Chemical structures of USFDA-approved non-nucleoside RT inhibitors (NNRTIs).</p>
Full article ">Figure 4
<p>Molecular arthritis. The structures of <span class="html-italic">apo</span> wild-type (WT) RT (left; PDB ID: 1DLO) and WT RT bound to efavirenz (right; PDB ID: 1FK9). Structures are represented by solvent-accessible surface area. The p51 subunit is colored in grey, and the p66 subunit is colored according to crystallographic B-factors, with blue being the least mobile and red the most mobile residues. EFV, efavirenz; NNRTIBP, NNRTI binding pocket.</p>
Full article ">
764 KiB  
Review
Biotechnological Applications of Transglutaminases
by Natalie M. Rachel and Joelle N. Pelletier
Biomolecules 2013, 3(4), 870-888; https://doi.org/10.3390/biom3040870 - 22 Oct 2013
Cited by 80 | Viewed by 11874
Abstract
In nature, transglutaminases catalyze the formation of amide bonds between proteins to form insoluble protein aggregates. This specific function has long been exploited in the food and textile industries as a protein cross-linking agent to alter the texture of meat, wool, and leather. [...] Read more.
In nature, transglutaminases catalyze the formation of amide bonds between proteins to form insoluble protein aggregates. This specific function has long been exploited in the food and textile industries as a protein cross-linking agent to alter the texture of meat, wool, and leather. In recent years, biotechnological applications of transglutaminases have come to light in areas ranging from material sciences to medicine. There has also been a substantial effort to further investigate the fundamentals of transglutaminases, as many of their characteristics that remain poorly understood. Those studies also work towards the goal of developing transglutaminases as more efficient catalysts. Progress in this area includes structural information and novel chemical and biological assays. Here, we review recent achievements in this area in order to illustrate the versatility of transglutaminases. Full article
(This article belongs to the Special Issue Enzymes and Their Biotechnological Applications)
Show Figures

Figure 1

Figure 1
<p>Amide bond formation catalyzed by TGase. Peptide- or protein-bound glutamines and lysines serve as substrates, releasing ammonia in the process.</p>
Full article ">Figure 2
<p>Crystal structure of MTG (PDB ID: 3IU0). The active site of the zymogen is covered (left) by an α-helix (gold), which is cleaved upon activation, exposing the active site cysteine residue (right, yellow spheres) that is critical for activity.</p>
Full article ">Figure 3
<p>Surface representation of MTG (PDB ID: 1UI4), illustrating active site residues investigated by mutagenesis (pink and orange regions) [<a href="#B48-biomolecules-03-00870" class="html-bibr">48</a>]. The active site cleft is indicated by an asterisk. Residues in orange, upon substitution to alanine, resulted in activity of 5% or less than the wild type, revealing their importance.</p>
Full article ">Figure 4
<p>Examples of assays used for detection of TGase activity. (<b>A</b>) Colorimetric and fluorescent product release activity assays. The hydroxamate assay (top) remains the standard method to determine and compare TGase activity. TG2 activity can also be quantified by the release of <span class="html-italic">p</span>-nitrophenol (PNP; λ<sub>max</sub> = 405 nm), umbelliferone (λ<sub>em</sub> = 465 nm), or by the formation of an anilide product (λ<sub>max</sub> = 278 nm) following conjugation with <span class="html-italic">N,N-</span>dimethyl-1,4-phenylenediamine (DMPDA). (<b>B</b>) Cartoon representation of the TG2 conformational FRET sensor. (<b>C</b>) <span class="html-italic">In vivo</span> activation of MTG allowing for in-cell assaying.</p>
Full article ">Figure 5
<p>General scheme for protein labeling using TGase. The protein of interest (P.O.I.) carries an accessible glutamine residue, for TGase-catalysed reaction with an amine-substituted fluorophore; alternatively, the P.O.I. carries a reactive lysine residue for reaction with a glutamine-modified fluorophore.</p>
Full article ">Figure 6
<p>Examples of TGases applied for visualization of biomacromolecules. (<b>A</b>) Locations of independently encoded 13-mer peptidyl loop K-tags on bacterial alkaline phosphatase. (<b>B</b>) MTG-aided enzymatic detection of nucleic acids. (<b>C</b>) The paramagnetic agent is cross-linked to a glutamine, generating the CEST effect. Magnetic resonance saturation is transferred to water following saturation of the amide proton.</p>
Full article ">
949 KiB  
Article
Variation in the Subcellular Localization and Protein Folding Activity among Arabidopsis thaliana Homologs of Protein Disulfide Isomerase
by Christen Y. L. Yuen, Kristie O. Matsumoto and David A. Christopher
Biomolecules 2013, 3(4), 848-869; https://doi.org/10.3390/biom3040848 - 21 Oct 2013
Cited by 25 | Viewed by 8891
Abstract
Protein disulfide isomerases (PDIs) catalyze the formation, breakage, and rearrangement of disulfide bonds to properly fold nascent polypeptides within the endoplasmic reticulum (ER). Classical animal and yeast PDIs possess two catalytic thioredoxin-like domains (a, a′) and two non-catalytic domains ( [...] Read more.
Protein disulfide isomerases (PDIs) catalyze the formation, breakage, and rearrangement of disulfide bonds to properly fold nascent polypeptides within the endoplasmic reticulum (ER). Classical animal and yeast PDIs possess two catalytic thioredoxin-like domains (a, a′) and two non-catalytic domains (b, b′), in the order a-b-b′-a′. The model plant, Arabidopsis thaliana, encodes 12 PDI-like proteins, six of which possess the classical PDI domain arrangement (AtPDI1 through AtPDI6). Three additional AtPDIs (AtPDI9, AtPDI10, AtPDI11) possess two thioredoxin domains, but without intervening b-b′ domains. C-terminal green fluorescent protein (GFP) fusions to each of the nine dual-thioredoxin PDI homologs localized predominantly to the ER lumen when transiently expressed in protoplasts. Additionally, expression of AtPDI9:GFP-KDEL and AtPDI10: GFP-KDDL was associated with the formation of ER bodies. AtPDI9, AtPDI10, and AtPDI11 mediated the oxidative folding of alkaline phosphatase when heterologously expressed in the Escherichia coli protein folding mutant, dsbA. However, only three classical AtPDIs (AtPDI2, AtPDI5, AtPDI6) functionally complemented dsbA. Interestingly, chemical inducers of the ER unfolded protein response were previously shown to upregulate most of the AtPDIs that complemented dsbA. The results indicate that Arabidopsis PDIs differ in their localization and protein folding activities to fulfill distinct molecular functions in the ER. Full article
(This article belongs to the Special Issue Protein Folding and Misfolding)
Show Figures

Figure 1

Figure 1
<p>Domain arrangements of PDIA1, PDIA6, and PDI-D homologs from <span class="html-italic">Arabidopsis</span>. Red boxes correspond to thioredoxin-like catalytic domains (<span class="html-italic">a</span>, <span class="html-italic">a′</span>, <span class="html-italic">a<sup>o</sup></span>), and grey boxes are non-catalytic thioredoxin-fold domains (<span class="html-italic">b, b′</span>). The signal peptides (sp) are represented as white boxes, and acidic regions (<span class="html-italic">c</span>) as yellow boxes. The <span class="html-italic">D</span> domain of PDI-D proteins is depicted as a blue box. The active site sequences of each catalytic domain are shown, with residues deviating from the typical CGHC motif underlined. The last four amino acids are shown at the end of each depicted protein.</p>
Full article ">Figure 2
<p><span class="html-italic">Arabidopsis</span> PDIA1 homologs fused to GFP localize to the ER. (<b>A</b>) PDI5:GFP-KDEL; (<b>B</b>) PDI6:GFP-KDEL; (<b>C</b>) PDI1:GFP-KDEL; (<b>D</b>) PDI2:GFP-KDEL; (<b>E</b>) PDI3:GFP-KDEL; (<b>F</b>) PDI4:GFP-KDEL. Each chimeric fusion was co-expressed in leaf protoplasts with the ER marker, ER:mCherry. GFP signal is shown in Column 1, mCherry signal in Column 2, and a merge of both signal patterns in Column 3. Chlorophyll autofluorescence is shown in Column 4. The white bars in Column 1 represent 10 µm.</p>
Full article ">Figure 3
<p><span class="html-italic">Arabidopsis</span> PDIA6 and PDI-D homologs fused to GFP localize to the ER. (<b>A</b>) PDI9:GFP-KDEL; (<b>B</b>) PDI10:GFP-KDDL; (<b>C</b>) PDI11:GFP; (<b>D</b>) unfused GFP(S65T) control. Each chimeric fusion was co-expressed in leaf protoplasts with the ER marker, ER:mCherry. GFP signal is shown in Column 1, mCherry signal in Column 2, and a merge of both signal patterns in Column 3. Chlorophyll autofluorescence is shown in Column 4. The white bars in Column 1 represent 10 µm.</p>
Full article ">Figure 4
<p>SESA1:mCherry partially localizes to punctate bodies induced by PDI9:GFP-KDEL or PDI10:GFP-KDDL expression. (<b>A</b>) PDI9:GFP-KDEL; (<b>B</b>) PDI10:GFP-KDDL; (<b>C</b>) PDI2:GFP-KDEL control (non-ER body-inducing); (<b>D</b>) unfused GFP(S56T) control. Each chimeric fusion was co-expressed in leaf protoplasts with the seed storage protein marker SESA1:mCherry. GFP signal is shown in Column 1, mCherry signal in Column 2, and a merge of both signal patterns in Column 3. Chlorophyll autofluorescence is shown in Column 4. Arrows indicate punctate bodies co-labeled by SESA1:mCherry and either (<b>A</b>) PDI9:GFP-KDEL or (<b>B</b>) PDI10:GFP-KDDL. The white bars in column 1 represent 10 µm.</p>
Full article ">Figure 5
<p>Alkaline phosphatase activity of <span class="html-italic">E. coli dsbA<sup>−</sup></span> cells expressing <span class="html-italic">Arabidopsis</span> homologs of PDIA1, PDIA6, and PDI-D. Measured alkaline phosphatase activities of wild-type strain RI89 (WT), <span class="html-italic">dsbA<sup>−</sup></span> strain RI90 (<span class="html-italic">dsbA<sup>−</sup></span>), and RI90 cells transformed with the pFLAG-CTS empty vector (EV) or constructs heterologously expressing the following <span class="html-italic">Arabidopsis</span> PDIs: (<b>A</b>) PDIA1 subfamily members AtPDI1 through AtPDI6; (<b>B</b>) PDIA6 subfamily members AtPDI9 and AtPDI10, and PDI-D subfamily member PDI11; (<b>C</b>) AtPDI1, AtPDI2, and modified variants AtPDI1<sub>M</sub> (<span class="html-italic">a</span>: CGHC) and AtPDI2<sub>M</sub> (<span class="html-italic">a</span>: CGAC); (<b>D</b>) AtPDI4 and modified variants AtPDI4<sub>M1</sub> (<span class="html-italic">a</span>: CARC), AtPDI4<sub>M2</sub> (<span class="html-italic">a</span>: CGHC), and AtPDI4<sub>M3</sub> (<span class="html-italic">a</span>: CGHC, <span class="html-italic">a′</span>: CGHC). Error bars represent standard deviations.</p>
Full article ">Figure 6
<p>Recombinant PDI expression in the protein folding mutant <span class="html-italic">dsbA</span><sup>−</sup> cells. Immunoblot of proteins extracted from mutant <span class="html-italic">E. coli</span> strain RI90 (<span class="html-italic">dsbA</span><sup>−</sup>), with or without constructs for the expression of recombinant PDIs. NV, no vector control; Vect, pFLAG-CTS empty vector control; PDI1, PDI2, PDI3, PDI4, PDI5 and PDI6, expression of the indicated PDI1A1 homolog; PDI9, expression of AtPDI9 (PDIA6 homolog); WT, <span class="html-italic">E. coli</span> strain RI89 (<span class="html-italic">dsbA</span><sup>+</sup>) control. Lanes denoted with a dash indicate empty lanes.</p>
Full article ">Figure 7
<p>Western blot analysis of PDI-GFP fusions expressed in <span class="html-italic">Arabidopsis</span> leaf protoplasts. Crude protein extracts were obtained from <span class="html-italic">Arabidopsis</span> leaf protoplasts transfected with constructs expressing GFP(S65T) alone (control) or fused to members of the <span class="html-italic">Arabidopsis</span> PDI family. The proteins were immobilized on a nitrocellulose membrane and GFP was detected using a rabbit anti-GFP primary antiserum, and an anti-rabbit horseradish peroxidase conjugated secondary antiserum. G, GFP(S65T); 1, PDI1-GFP-KDEL; 2, PDI2-GFP-KDEL; 3, PDI3-GFP-KDEL; 4, PDI4-GFP-KDEL; 5, PDI5-GFP-KDEL; 6, PDI6-GFP-KDEL; PDI9-GFP-KDEL; 10, PDI10-GFP-KDDL; 11, PDI11-GFP.</p>
Full article ">
674 KiB  
Review
Biocatalysis for Biobased Chemicals
by Rubén De Regil and Georgina Sandoval
Biomolecules 2013, 3(4), 812-847; https://doi.org/10.3390/biom3040812 - 17 Oct 2013
Cited by 42 | Viewed by 14467
Abstract
The design and development of greener processes that are safe and friendly is an irreversible trend that is driven by sustainable and economic issues. The use of Biocatalysis as part of a manufacturing process fits well in this trend as enzymes are themselves [...] Read more.
The design and development of greener processes that are safe and friendly is an irreversible trend that is driven by sustainable and economic issues. The use of Biocatalysis as part of a manufacturing process fits well in this trend as enzymes are themselves biodegradable, require mild conditions to work and are highly specific and well suited to carry out complex reactions in a simple way. The growth of computational capabilities in the last decades has allowed Biocatalysis to develop sophisticated tools to understand better enzymatic phenomena and to have the power to control not only process conditions but also the enzyme’s own nature. Nowadays, Biocatalysis is behind some important products in the pharmaceutical, cosmetic, food and bulk chemicals industry. In this review we want to present some of the most representative examples of industrial chemicals produced in vitro through enzymatic catalysis. Full article
(This article belongs to the Special Issue Enzymes and Their Biotechnological Applications)
Show Figures

Figure 1

Figure 1
<p>Enzymatic synthesis of Difructose Anhydride III (DFA III) from Inulin using enzyme inulase II, an inulin fructo-transferase.</p>
Full article ">Figure 2
<p>Simplified enzymatic reaction to produce Galacto-oligosaccharides (GOS) from lactose using β-galactosidase. Subscript <span class="html-italic">n</span> may range from 0–6 though most native enzymes produce GOS with <span class="html-italic">n</span> between 1–2.</p>
Full article ">Figure 3
<p>Simplified enzymatic reactions used commercially to obtain Fructo-oligosaccharides (FOS).</p>
Full article ">Figure 4
<p>The main two approaches to synthesize MLM Structured Lipids enzymatically: (<b>A</b>) Transesterification and (<b>B</b>) Acidolysis. Subproducts like LML Structured lipids may also be obtained, but only the product of interest is shown. L = Long Chain; M = Medium Chain; MCFA = Medium Chain Fatty Acid; MCFAEt = Medium Chain Fatty Acid Ester (see also [<a href="#B48-biomolecules-03-00812" class="html-bibr">48</a>]).</p>
Full article ">Figure 5
<p>Enzymatic reaction of transesterification of oils with alcohol to produce fatty acid esters (biodiesel).</p>
Full article ">Figure 6
<p>Mechanism of phenol polymer formation (see also [<a href="#B104-biomolecules-03-00812" class="html-bibr">104</a>]).</p>
Full article ">Figure 7
<p>Enzymatic conversion of acrylonitrile to acrylamide by means of a nitrile hydratase.</p>
Full article ">Figure 8
<p>Chemoenzymatic route to obtain nicotinamide developed by Lonza [<a href="#B209-biomolecules-03-00812" class="html-bibr">209</a>,<a href="#B210-biomolecules-03-00812" class="html-bibr">210</a>].</p>
Full article ">Figure 9
<p>Biocatalytic synthesis of thymidine using a Purine Nucleoside Phosphorilase (PNPase) and the biocatalytical removal of hypoxantine with a Xanthine Oxidase (XAO) to obtain uric acid. Compounds: 1: 2’deoxyinosine; 2: Thymine; 3: Thymidine; 4: Hypoxanthine; 5: Uric acid (see also [<a href="#B233-biomolecules-03-00812" class="html-bibr">233</a>]).</p>
Full article ">
1286 KiB  
Review
Improvement of Biocatalysts for Industrial and Environmental Purposes by Saturation Mutagenesis
by Francesca Valetti and Gianfranco Gilardi
Biomolecules 2013, 3(4), 778-811; https://doi.org/10.3390/biom3040778 - 8 Oct 2013
Cited by 19 | Viewed by 16891
Abstract
Laboratory evolution techniques are becoming increasingly widespread among protein engineers for the development of novel and designed biocatalysts. The palette of different approaches ranges from complete randomized strategies to rational and structure-guided mutagenesis, with a wide variety of costs, impacts, drawbacks and relevance [...] Read more.
Laboratory evolution techniques are becoming increasingly widespread among protein engineers for the development of novel and designed biocatalysts. The palette of different approaches ranges from complete randomized strategies to rational and structure-guided mutagenesis, with a wide variety of costs, impacts, drawbacks and relevance to biotechnology. A technique that convincingly compromises the extremes of fully randomized vs. rational mutagenesis, with a high benefit/cost ratio, is saturation mutagenesis. Here we will present and discuss this approach in its many facets, also tackling the issue of randomization, statistical evaluation of library completeness and throughput efficiency of screening methods. Successful recent applications covering different classes of enzymes will be presented referring to the literature and to research lines pursued in our group. The focus is put on saturation mutagenesis as a tool for designing novel biocatalysts specifically relevant to production of fine chemicals for improving bulk enzymes for industry and engineering technical enzymes involved in treatment of waste, detoxification and production of clean energy from renewable sources. Full article
(This article belongs to the Special Issue Enzymes and Their Biotechnological Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of site saturation mutagenesis approach following the QuikChange<sup>TM</sup> kit.</p>
Full article ">Figure 2
<p>Scheme of iterative saturation mutagenesis showing the branching process and highlighting the productive pathway (in green), non-productive mutants that stop the process simplifying the screening procedure (in red). Highlighted in yellow are mutants produced with moderate to low improvement that can be discarded or reconsidered for further processing in a second phase.</p>
Full article ">Figure 3
<p>Schematic of PFunkel mutagenesis strategy (adapted from Firnberg <span class="html-italic">et al</span>. [<a href="#B44-biomolecules-03-00778" class="html-bibr">44</a>]).</p>
Full article ">Figure 4
<p>The 4-step strategy for the simultaneous saturation of 5 independent codons by OmniChange (adapted from Dennig <span class="html-italic">et al</span>. [<a href="#B45-biomolecules-03-00778" class="html-bibr">45</a>]).</p>
Full article ">Figure 5
<p>DNA sequencing of the three libraries produced for evaluation of the randomization efficiency on selected position in hydrogenase gene: the targeted position is properly randomized for NNK in library A and C (K either a T or a G), while only partial degeneration is present in library B.</p>
Full article ">Figure 6
<p>(<b>a</b>) Scheme of the principle of on-colonies activity test for a [FeFe] hydrogenase [<a href="#B55-biomolecules-03-00778" class="html-bibr">55</a>]; (<b>b</b>) Example of the screening results.</p>
Full article ">Figure 7
<p>Scheme of the structure of <span class="html-italic">P. aeruginosa</span> lipase active site pocket (PDB: 1EX9) with the targeted sites (library A: Met16/Leu17 in red; library B: Leu118/Ile 121 in orange; library C: Leu131/Val 135 in yellow; library D: Leu159/Leu162 in green; and library E: Leu231/Val 232 in cyan). Ser82, Asp229, and His251 (in violet) represent the catalytic triad. A substrate analogue (RC-(RP,SP)-1,2-dioctylcarbamoyl-glycero-3-O-octylphosphonate) covalently bound to Ser 82 is shown in blue.</p>
Full article ">Figure 8
<p>Structure of P450 BM3 heme domain (PDB: 2HPD) showing the target sites A (Arg47, Thr49, Tyr51) in green and B (Val78, Ala82) in blue. Heme is shown in red, the Fe coordinating Cys 400 is in magenta.</p>
Full article ">Figure 9
<p>Scheme of the active site of PAMO (PDB: 1W4X) with targeted residues Pro440, Pro437, Gln93 and Pro94 (in black). FAD is shown in orange; Arg 337, involved in catalysis, is shown in blue.</p>
Full article ">Figure 10
<p>(<b>a</b>) Structure of active site of catechol 1,2 dioxygenase highlighting the residue that define the active site pocket (PDB file from crystallographic structure in [<a href="#B117-biomolecules-03-00778" class="html-bibr">117</a>]); (<b>b</b>) The effect of reshaping by mutagenesis and SSM on model of substrate/pocket interaction; (<b>c</b>) The list of identified and characterized mutants for SSM on position 72 are reported in the table (related to studies published in [<a href="#B114-biomolecules-03-00778" class="html-bibr">114</a>]).</p>
Full article ">Figure 11
<p>Model of CaHydA structure illustrating C298 (left) and replacement at 298 position with aspartic acid (right) (adapted from Morra <span class="html-italic">et al</span>. [<a href="#B55-biomolecules-03-00778" class="html-bibr">55</a>]).</p>
Full article ">
3082 KiB  
Review
Biocatalytic Synthesis of Chiral Alcohols and Amino Acids for Development of Pharmaceuticals
by Ramesh N. Patel
Biomolecules 2013, 3(4), 741-777; https://doi.org/10.3390/biom3040741 - 2 Oct 2013
Cited by 125 | Viewed by 16993
Abstract
Chirality is a key factor in the safety and efficacy of many drug products and thus the production of single enantiomers of drug intermediates and drugs has become increasingly important in the pharmaceutical industry. There has been an increasing awareness of the enormous [...] Read more.
Chirality is a key factor in the safety and efficacy of many drug products and thus the production of single enantiomers of drug intermediates and drugs has become increasingly important in the pharmaceutical industry. There has been an increasing awareness of the enormous potential of microorganisms and enzymes derived there from for the transformation of synthetic chemicals with high chemo-, regio- and enatioselectivities. In this article, biocatalytic processes are described for the synthesis of chiral alcohols and unntural aminoacids for pharmaceuticals. Full article
(This article belongs to the Special Issue Enzymes and Their Biotechnological Applications)
Show Figures

Figure 1

Figure 1
<p>Hydroxy buspirone (antianxiety drug): Enzymatic preparation of 6-hydroxybuspirone.</p>
Full article ">Figure 2
<p>Cholesterol lowering agents: Enzymatic preparation of (3<span class="html-italic">S</span>,5<span class="html-italic">R</span>)-dihydroxy-6- (benzyloxy) hexanoic acid, ethyl ester.</p>
Full article ">Figure 3
<p>Atorvastatin: Enzymatic preparation of (<span class="html-italic">R</span>)-4-cyano-3-hydroxybutyrate.</p>
Full article ">Figure 4
<p>Chloesterol lowering agents: Preparation of (<span class="html-italic">S</span>)-4-chloro-3-hydroxybutanoic acid methyl ester.</p>
Full article ">Figure 5
<p>Rhinovirus protease inhibitor: Enzymatic process for the preparation of (<span class="html-italic">R</span>)-3-(4-fluorophenyl)-2-hydroxy propionic acid.</p>
Full article ">Figure 6
<p>Atazanavir (antiviral agent): Enzymatic reparation of (1<span class="html-italic">S</span>,2<span class="html-italic">R</span>)-[3-chloro-2-hydroxy-1-(phenylmethyl) propyl]-carbamic acid,1,1-dimethyl-ethyl ester.</p>
Full article ">Figure 7
<p>Enzymatic reduction process for synthesis (<span class="html-italic">S</span>)-alcohol <b>27</b> for Montelukast intermediate.</p>
Full article ">Figure 8
<p>Anticancer drug: Enzymatic preparation of C-13 paclitaxel side-chain synthon.</p>
Full article ">Figure 9
<p>Antipsychotic drug: Enzymatic reduction of 1-(4-fluorophenyl)4-[4-(5-fluoro-2-pyrimidinyl)1-piperazinyl]-1-butanone.</p>
Full article ">Figure 10
<p>Retinoic acid receptor agonist: Enzymatic preparation of 2-(<span class="html-italic">R</span>)-hydroxy-2-(1',2',3',4'-tetrahydro-1',1',4',4'-tetramethyl-6'-naphthalenyl)acetate.</p>
Full article ">Figure 11
<p>Anti-Alzheimer’s drugs: Enzymatic reduction of 5-oxohexanoate and 5-oxohexanenitrile.</p>
Full article ">Figure 12
<p>(<b>A</b>) Anti-Alzheimer’s drugs: Enantioselective microbial reduction of substituted acetophenone; (<b>B</b>) Enantioselective microbial reduction of methyl-4-(2'-acetyl-5'-fluorophenyl) butanoates.</p>
Full article ">Figure 13
<p>Anticancer drug: Enzymatic preparation of (<span class="html-italic">S</span>)-2-chloro-1-(3-chlorophenyl)ethanol.</p>
Full article ">Figure 14
<p>Thrombin inhibitor: Enzymatic preparation of (<span class="html-italic">R</span>)-2-Hydroxy-3,3-dimethylbutanoic acid.</p>
Full article ">Figure 15
<p>Endothelin receptor antagonist: Enantioselective microbial reduction of keto ester and chloroketone.</p>
Full article ">Figure 16
<p>Calcium channel blocker: Preparation of [(3<span class="html-italic">R</span>-<span class="html-italic">cis</span>)-1,3,4,5-tetrahydro-3-hydroxy-4-(4-methoxyphenyl)-6-(trifluromethyl)-2<span class="html-italic">H</span>-1-benzazepin-2-one].</p>
Full article ">Figure 17
<p>β3-Receptor agonist: Reduction of 4-benzyloxy-3-methanesulfonylamino-2-bromo-acetophenone.</p>
Full article ">Figure 18
<p>Penem and carbapenem: Enzymatic preparation of (<span class="html-italic">R</span>)-1,3-butanediol and (<span class="html-italic">R</span>)-4-chloro-3-hydroxybutonoate.</p>
Full article ">Figure 19
<p>Integrin receptor agonist: Enzymatic preparation of (<span class="html-italic">R</span>)-allylic alcohol.</p>
Full article ">Figure 20
<p>NK1 receptor antagonists: Enzymatic synthesis of (<span class="html-italic">S</span>)-3,5-bistrifluoromethylphenyl ethanol.</p>
Full article ">Figure 21
<p>Tigemonam: Enzymatic synthesis of (<span class="html-italic">S</span>)-β-hydroxyvaline.</p>
Full article ">Figure 22
<p>Atazanavir (anti-viral agent): Enzymatic synthesis of (<span class="html-italic">S</span>)-tertiary-leucine.</p>
Full article ">Figure 23
<p>Vanlev: Enzymatic synthesis of (<span class="html-italic">S</span>)-6-hydroxynorleucine by reductive amination.</p>
Full article ">Figure 24
<p>Vanlev: Enzymatic conversion of racemic 6-hydroxy norleucine to (<span class="html-italic">S</span>)-6-hydroxymorleucine.</p>
Full article ">Figure 25
<p>Vanlev: Enzymatic synthesis of allysine ethylene acetal.</p>
Full article ">Figure 26
<p>Saxagliptin: Enzymatic reductive amination of 2-(3-hydroxy-1-adamantyl)-2-oxoethanoic acid.</p>
Full article ">Figure 27
<p>Enzymatic synthesis of (<span class="html-italic">S</span>)-neopentylglycine.</p>
Full article ">Figure 28
<p>Glucogen like peptide: The (<span class="html-italic">S</span>)-amino-3-[3-{6-(2-methylphenyl)}pyridyl]-propionic acid.</p>
Full article ">Figure 29
<p>Thrombin inhibitor inogatran: Enzymatic synthesis of (<span class="html-italic">R</span>)-cyclohexylalanine.</p>
Full article ">Figure 30
<p>Calcitonin gene-related peptide receptors (antimigraine drugs): Enzymatic preparation of (<span class="html-italic">R</span>)-2-amino-3-(7-methyl-1<span class="html-italic">H</span>-indazol-5-yl)propanoic acid.</p>
Full article ">Figure 31
<p>Corticotropin releasing factor (CRF)-1 receptor antagonist: Enzymatic synthesis of (<span class="html-italic">R</span>)-1-cyclopropylethylamine and (<span class="html-italic">R</span>)-<span class="html-italic">sec</span>-butylamine.</p>
Full article ">
396 KiB  
Article
Isolation, NMR Spectral Analysis and Hydrolysis Studies of a Hepta Pyranosyl Diterpene Glycoside from Stevia rebaudiana Bertoni
by Venkata Sai Prakash Chaturvedula, Steven Chen, Oliver Yu and Guohong Mao
Biomolecules 2013, 3(4), 733-740; https://doi.org/10.3390/biom3040733 - 30 Sep 2013
Cited by 7 | Viewed by 6615
Abstract
From the commercial extract of the leaves of Stevia rebaudiana Bertoni, a minor steviol glycoside, 13-[(2-O-β-D-glucopyranosyl-3-O-β-D-glucopyranosyl-β-D-glucopyranosyl)oxy] ent-kaur-16-en-19-oic acid-[(2-O-(3-O-β-D-glucopyranosyl-α-L-rhamnopyranosyl)-3-O-β-D-glucopyranosyl-β-D-glucopyranosyl) ester] (1); also known as rebaudioside O having seven sugar units has [...] Read more.
From the commercial extract of the leaves of Stevia rebaudiana Bertoni, a minor steviol glycoside, 13-[(2-O-β-D-glucopyranosyl-3-O-β-D-glucopyranosyl-β-D-glucopyranosyl)oxy] ent-kaur-16-en-19-oic acid-[(2-O-(3-O-β-D-glucopyranosyl-α-L-rhamnopyranosyl)-3-O-β-D-glucopyranosyl-β-D-glucopyranosyl) ester] (1); also known as rebaudioside O having seven sugar units has been isolated. Its structural characterization has been achieved by the extensive 1D (1H and 13C), and 2D NMR (COSY, HMQC, HMBC) as well as mass spectral data. Further, hydrolysis studies were performed on rebaudioside O using acid and enzymatic methods to identify aglycone and sugar residues in its structure as well as their configurations. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of rebaudioside O (<b>1</b>), rebaudioside N (<b>2</b>), and steviol (<b>3</b>).</p>
Full article ">Figure 2
<p>Key COSY and HMBC correlations of rebaudioside O (<b>1</b>).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop