[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (488)

Search Parameters:
Keywords = substrate recognition

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 3742 KiB  
Article
Comparison of Survivin Determination by Surface-Enhanced Fluorescence and Raman Spectroscopy on Nanostructured Silver Substrates
by Georgia Geka, Anastasia Kanioura, Ioannis Kochylas, Vlassis Likodimos, Spiros Gardelis, Anastasios Dimitriou, Nikolaos Papanikolaou, Anastasios Economou, Sotirios Kakabakos and Panagiota Petrou
Biosensors 2024, 14(10), 479; https://doi.org/10.3390/bios14100479 - 6 Oct 2024
Viewed by 601
Abstract
Survivin belongs to a family of proteins that promote cellular proliferation and inhibit cellular apoptosis. Its overexpression in various cancer types has led to its recognition as an important marker for cancer diagnosis and treatment. In this work, we compare two approaches for [...] Read more.
Survivin belongs to a family of proteins that promote cellular proliferation and inhibit cellular apoptosis. Its overexpression in various cancer types has led to its recognition as an important marker for cancer diagnosis and treatment. In this work, we compare two approaches for the immunochemical detection of survivin through surface-enhanced fluorescence or Raman spectroscopy using surfaces with nanowires decorated with silver nanoparticles in the form of dendrites or aggregates as immunoassays substrates. In both substrates, a two-step non-competitive immunoassay was developed using a pair of specific monoclonal antibodies, one for detection and the other for capture. The detection antibody was biotinylated and combined with streptavidin labeled with rhodamine for the detection of surface-enhanced fluorescence, while, for the detection via Raman spectroscopy, streptavidin labeled with peroxidase was used and the signal was obtained after the application of 3,3′,5,5′-tetramethylbenzidine (TMB) precipitating substrate. It was found that the substrate with the silver dendrites provided higher fluorescence signal intensity compared to the substrate with the silver aggregates, while the opposite was observed for the Raman signal. Thus, the best substrate was used for each detection method. A detection limit of 12.5 pg/mL was achieved with both detection approaches along with a linear dynamic range up to 500 pg/mL, enabling survivin determination in human serum samples from both healthy and ovarian cancer patients for cancer diagnosis and monitoring purposes. Full article
(This article belongs to the Special Issue Noble Metal Nanoparticle-Based Nanoplatforms for Biosensors)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the immunoassay for survivin determination with SEF (<b>a</b>) or SERS (<b>b</b>) and of the reagents used in both assays (<b>c</b>).</p>
Full article ">Figure 2
<p>Characteristic cross and top view SEM images of Si nanowires with height of approximately 250 nm decorated with 800 nm-long Ag dendrites (<b>a</b>,<b>b</b>) or Si nanowires with height of approximately 500 nm decorated with approximately 150 nm-long Ag aggregates (<b>c</b>,<b>d</b>). The cross-view images (<b>a</b>,<b>c</b>) have been received with a higher magnification to show in detail the Ag nanoparticles’ structure whereas the top view images (<b>b</b>,<b>d</b>) with a lower magnification to demonstrate the distribution of these structures across the substrate surface.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) SEF spectra obtained from nanostructured Ag/Si substrates with Ag dendrites following a one-step (<b>a</b>) or a two-step survivin immunoassay (<b>b</b>) for the zero calibrator (black line) and calibrators containing 50 (red line) and 500 pg/mL (blue line). (<b>c</b>,<b>d</b>) SERS spectra obtained from nanostructured Ag/Si substrates with Ag aggregates following a one-step (<b>c</b>) or a two-step survivin immunoassay (<b>d</b>) for the zero calibrator (black line) and calibrators containing 50 (red line) and 500 pg/mL (green line).</p>
Full article ">Figure 4
<p>(<b>a</b>) SEF signal values corresponding to zero calibrator (plain orange and green columns) and a calibrator containing 50 pg/mL of survivin (striped orange and green columns) obtained using the detection antibody at concentrations of 1.25 (orange columns) and 2.5 μg/mL (green columns). Each point is the mean of three samples ± SD. (<b>b</b>) Survivin calibration curves obtained from nanostructured Ag/Si substrates with Ag dendrites for duration of each immunoassay step equal to 0.5 (squares), 1 (circles), or 2 h (triangles). Each point is the mean of three samples ± SD.</p>
Full article ">Figure 5
<p>(<b>a</b>) Raman signals corresponding to peak at 1607 cm<sup>−1</sup> obtained for the zero calibrator (squares) or calibrators containing 25 (circles) or 200 pg/mL (triangles) of survivin with respect to the streptavidin-peroxidase concentration. The incubation with the precipitating TMB substrate was 15 min. Each point is the mean of three samples ± SD. (<b>b</b>) Raman signals corresponding to peak at 1607 cm<sup>−1</sup> obtained for the zero calibrator (squares) or calibrators containing 25 (circles) or 200 pg/mL (triangles) of survivin with respect to the duration of incubation with the precipitating TMB substrate. The streptavidin-peroxidase concentration was 25 ng/mL. Each point is the mean of three samples ± SD.</p>
Full article ">Figure 6
<p>(<b>a</b>) SEF spectra received from Ag/Si substrates with dendrites for survivin calibrators from 0 to 500 pg/mL in assay buffer. (<b>b</b>) Survivin SEF calibration curve. (<b>c</b>) SERS spectra received from Ag/Si substrates with aggregates for survivin calibrators from 0 to 500 pg/mL in assay buffer. (<b>d</b>) Survivin SERS calibration curve. Each point is the mean value of 10 measurements from 3 replicate samples ± 3SD.</p>
Full article ">
13 pages, 2510 KiB  
Article
Sandwich-Type Electrochemical Aptasensor with Supramolecular Architecture for Prostate-Specific Antigen
by Anabel Villalonga, Raúl Díaz, Irene Ojeda, Alfredo Sánchez, Beatriz Mayol, Paloma Martínez-Ruiz, Reynaldo Villalonga and Diana Vilela
Molecules 2024, 29(19), 4714; https://doi.org/10.3390/molecules29194714 - 5 Oct 2024
Viewed by 403
Abstract
A novel sandwich-type electrochemical aptasensor based on supramolecularly immobilized affinity bioreceptor was prepared via host–guest interactions. This method utilizes an adamantane-modified, target-responsive hairpin DNA aptamer as a capture molecular receptor, along with a perthiolated β-cyclodextrin (CD) covalently attached to a gold-modified electrode surface [...] Read more.
A novel sandwich-type electrochemical aptasensor based on supramolecularly immobilized affinity bioreceptor was prepared via host–guest interactions. This method utilizes an adamantane-modified, target-responsive hairpin DNA aptamer as a capture molecular receptor, along with a perthiolated β-cyclodextrin (CD) covalently attached to a gold-modified electrode surface as the transduction element. The proposed sensing strategy employed an enzyme-modified aptamer as the signalling element to develop a sandwich-type aptasensor for detecting prostate-specific antigen (PSA). To achieve this, screen-printed carbon electrodes (SPCEs) with electrodeposited reduced graphene oxide (RGO) and gold nanoferns (AuNFs) were modified with the CD derivative to subsequently anchor the adamantane-modified anti-PSA aptamer via supramolecular associations. The sensing mechanism involves the affinity recognition of PSA molecules on the aptamer-enriched electrode surface, followed by the binding of an anti-PSA aptamer–horseradish peroxidase complex as a labelling element. This sandwich-type arrangement produces an analytical signal upon the addition of H2O2 and hydroquinone as enzyme substrates. The aptasensor successfully detected the biomarker within a concentration range of 0.5 ng/mL to 50 ng/mL, exhibiting high selectivity and a detection limit of 0.11 ng/mL in PBS. Full article
(This article belongs to the Special Issue Nano-Functional Materials for Sensor Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representative FE-SEM images of bare SPCE (<b>A</b>), rGO/SPCE (<b>B</b>), AuNFs/rGO/SPCE (<b>C</b>), CD–AuNFs/rGO/SPCE (<b>D</b>), Apt–ADA/CD–AuNFs/rGO/SPCE (<b>E</b>), MCH/Apt–ADA/CD–AuNFs/rGO/SPCE (<b>F</b>).</p>
Full article ">Figure 2
<p>Cyclic voltammograms (<b>A</b>) and Nyquist plots (<b>B</b>) of SPCE before (a) and after sequential modification with rGO (b), AuNFs (c), CD (d), Apt–ADA (e), and MCH (f). Cyclic voltammograms (<b>C</b>) and Nyquist plots (<b>D</b>) of the sensor (a) and further incubation with PSA (b) and Apt–HRP (c). Measured in 0.1 M KCl solution containing 5 mM K<sub>3</sub>[Fe(CN)<sub>6</sub>]/K<sub>4</sub>[Fe(CN)<sub>6</sub>] (1:1), (<b>A</b>,<b>C</b>) scan rate  =  50 mV·s<sup>−1</sup>. Conditions for EIS (<b>B</b>,<b>D</b>): frequency range of 0.01 to 106 Hz at a fixed potential.</p>
Full article ">Figure 3
<p>Nyquist plots of Apt–ADA/CD-AuNFs/rGO/SPCE electrode (<b>A</b>) before (a) and after sequential incubation with NaOH (b), MCH (c), and PSA (d). Nyquist plots of Apt–ADA/CD–AuNFs/rGO/SPCE electrode (<b>B</b>) before (a) and after sequential incubation with ADA-COOH in NaOH (b), MCH (c) and PSA (d). Measured in 0.100 M KCl solution containing 5 mM K<sub>3</sub>[Fe(CN)<sub>6</sub>]/K<sub>4</sub>[Fe(CN)<sub>6</sub>] (1:1).</p>
Full article ">Figure 4
<p>Amperometric responses (<b>A</b>) and relative amperometric analytical signal (<b>B</b>) of the aptasensor, previously incubated with NaOH (grey) or ADA-COOH in NaOH (red), towards PSA.</p>
Full article ">Figure 5
<p>(<b>A</b>) Calibration plot for the aptasensor towards PSA in 0.1 M sodium phosphate buffer, pH 7.0. (<b>B</b>) Relative amperometric response of the aptasensor toward 50 ng·mL<sup>−1</sup> PSA and 100 ng·mL<sup>−1</sup> of CEA, HSA, IgG, and TBA, respectively. (<b>C</b>) Relative amperometric response of the aptasensor toward PSA and mixtures with other potential interfering proteins at the same concentration cited above.</p>
Full article ">Scheme 1
<p>Schematic representation of the processes involved in the assembly (<b>A</b>) and method of use (<b>B</b>) of the aptasensor for PSA.</p>
Full article ">
16 pages, 7031 KiB  
Article
Structural Insights and Catalytic Mechanism of 3-Hydroxybutyryl-CoA Dehydrogenase from Faecalibacterium Prausnitzii A2-165
by Jaewon Yang, Hyung Jin Jeon, Seonha Park, Junga Park, Seonhye Jang, Byeongmin Shin, Kyuhyeon Bang, Hye-Jin Kim Hawkes, Sungha Park, Sulhee Kim and Kwang Yeon Hwang
Int. J. Mol. Sci. 2024, 25(19), 10711; https://doi.org/10.3390/ijms251910711 - 5 Oct 2024
Viewed by 464
Abstract
Atopic dermatitis (AD) is characterized by a T-helper cell type 2 (Th2) inflammatory response leading to skin damage with erythema and edema. Comparative fecal sample analysis has uncovered a strong correlation between AD and Faecalibacterium prausnitzii strain A2-165, specifically associated with butyrate production. [...] Read more.
Atopic dermatitis (AD) is characterized by a T-helper cell type 2 (Th2) inflammatory response leading to skin damage with erythema and edema. Comparative fecal sample analysis has uncovered a strong correlation between AD and Faecalibacterium prausnitzii strain A2-165, specifically associated with butyrate production. Therefore, understanding the functional mechanisms of crucial enzymes in the butyrate pathway, such as 3-hydroxybutyryl-CoA dehydrogenase of A2-165 (A2HBD), is imperative. Here, we have successfully elucidated the three-dimensional structure of A2HBD in complex with acetoacetyl-CoA and NAD+ at a resolution of 2.2Å using the PAL-11C beamline (third generation). Additionally, X-ray data of A2HBD in complex with acetoacetyl-CoA at a resolution of 1.9 Å were collected at PAL-XFEL (fourth generation) utilizing Serial Femtosecond Crystallography (SFX). The monomeric structure of A2HBD consists of two domains, N-terminal and C-terminal, with cofactor binding occurring at the N-terminal domain, while the C-terminal domain facilitates dimerization. Our findings elucidate the binding mode of NAD+ to A2HBD. Upon acetoacetyl-CoA binding, the crystal structure revealed a significant conformational change in the Clamp-roof domain (root-mean-square deviation of 2.202 Å). Notably, residue R143 plays a critical role in capturing the adenine phosphate ring, underlining its significance in substrate recognition and catalytic activity. The binding mode of acetoacetyl-CoA was also clarified, indicating its lower stability compared to NAD+. Furthermore, the conformational change of hydrophobic residues near the catalytic cavity upon substrate binding resulted in cavity shrinkage from an open to closed conformation. This study confirms the conformational changes of catalytic triads involved in the catalytic reaction and presents a proposed mechanism for substrate reduction based on structural observations. Full article
(This article belongs to the Special Issue Structural Dynamics of Macromolecules)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Overall structure of apo-A2HBD: Each subunit of hexameric A2HBD is distinguished with different colors (<b>b</b>) SEC-MALS of apo-3-Hydroxybutyryl-CoA (A2HBD); sample was measured at 2.018 mg/mL concentration. The protein sample was separated on a size exclusion chromatography (SEC) column. The horizontal line marked as red in yellow circle. represents the measured molar mass. The peak (black) represents the light scattering.</p>
Full article ">Figure 2
<p>The overall structure of A2HBD in complex with NAD<sup>+</sup> and acetoacetyl-CoA. (<b>a</b>) The merged hexamer of the A2HBD trimer and the symmetric subunit are shown as a cartoon. The trimeric asymmetric unit of the tertiary complex is shown as a colored cartoon. Subunits A and C are labeled as blue and cyan, respectively. The B subunit of the tertiary complex is shown in yellow. The bound ligand NAD<sup>+</sup> and acetoacetyl-CoA are shown in magenta and white, respectively, in a stick model. The symmetric structure of A2HBD is shown in white. (<b>b</b>) Comparison of the cleft angle in subunit dimers. Non-bound subunits A and C dimer are shown as cartoons. Subunits A and C are colored blue and cyan, respectively. (<b>c</b>) The merged dimer of ligand-bound subunit B and its symmetric unit are shown as cartoons with yellow and white, respectively. The cleft angle is defined by Lys56 at the edge of α2 of the two subunits and His220 at the center of the dimer. The tilted cleft angle is further narrowed by the addition of ligands. Ligands of NAD<sup>+</sup> and acetoacetyl-CoA are shown as stick models with magenta and white, respectively.</p>
Full article ">Figure 3
<p>The overall structure of A2HBD in complex with NAD<sup>+</sup> and acetoacetyl-CoA. (<b>a</b>) Monomer structure of A2HBD complexed with NAD<sup>+</sup> and acetoacetyl-CoA. The tertiary monomeric structure is depicted in a cartoon representation, with the N-terminal domain highlighted in green and the C-terminal domain in cyan. The presence of bound NAD<sup>+</sup> and acetoacetyl-CoA is illustrated using a stick model, with NAD<sup>+</sup> represented in magenta and acetoacetyl-CoA in white. The regions responsible for substrate binding, depicted as the clamp-roof and clamp-base domains, are indicated by a dotted circle in black and are appropriately labeled. (<b>b</b>) Crystal structure of A2HBD in complex with NAD<sup>+</sup>. (left) The “open-book” view of A2HBD monomer is shown in both the cartoon and surface diagram. Three sub-domains and two domains are labeled. The stick diagram of NAD<sup>+</sup> is colored magenta. (right) An enlarged view of the cofactor-binding pocket. NAD<sup>+</sup> is shown as a stick and A2HBD is shown as a surface. The subsites (S1, S2, S3) of A2HBD are labeled. (<b>c</b>) The detailed interaction between NAD<sup>+</sup> and A2HBD. The G-x-G-x-x-G motif is marked as yellow. Hydrogen bonding interactions are indicated as black dashed lines. (<b>d</b>) Schematic interactions between NAD<sup>+</sup> and A2HBD. Hydrogen bonding interactions are indicated as black dashed lines, whereas hydrophobic interactions are not indicated in lines.</p>
Full article ">Figure 4
<p>Dimerization interface of A2HBD (<b>a</b>) The representative snapshot of the crystal structure of the A2HBD dimer is shown in the cartoon (left) and the surface (right) model. In the cartoon diagram, the subunits A and C are denoted by blue and cyan colors, respectively. They are arranged in a “tail-to-tail” configuration, facilitated by interactions between their C-terminal domains. The surface diagram highlights only the secondary structures relevant to dimerization. (<b>b</b>) An “open-book” view of the dimerization interface between subunits A and C. The contact sites of the dimerization interface are indicated with red, green, and blue arrows; and dotted figures. (<b>c</b>) Core residues involved in dimerization interface. The combined ribbon and stick model depicts the electrostatic interactions (left) and hydrophobic interactions (right) between the secondary structure elements that contribute to the dimerization interface of the two subunits.</p>
Full article ">Figure 5
<p>Domain shifting of A2HBD upon acetoacetyl-CoA binding. (<b>a</b>) Superimposition of NAD<sup>+</sup> complex A2HBD (white) and tertiary complex (green and cyan). Clamp-roof domain movement (α2 to α2′) is described in (<b>a</b>) dotted box. NAD<sup>+</sup> in cofactor complex, which is transparent, relocated to clear NAD<sup>+</sup> (magenta) upon substrate binding. (<b>b</b>) The detailed cartoon diagram of clamp-roof domain shifting with (<b>a</b>) dihedral angle of 11 degrees upon substrate binding. (<b>c</b>) Surface diagram of open-to-closed conformation followed by substrate binding. (<b>d</b>,<b>e</b>) The open-close conformational change in residues of the catalytic cavity. Amino acid residues involved in forming catalytic cavities are shown as a stick model with a yellow color. “Catalytic cavity” is presented with a dotted circle. Upon substrate binding, the size of the cavity decreased by about 55%. The cation-π interaction (Asn139-Phe230) is shown as dotted red lines.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) The open-closed conformational change of interactions within catalytic residues in A2HBD upon substrate binding. In both panels, A2HBD structures are shown as gray-colored cartoons. The catalytic residues (Ser117, His136, Glu148) and NAD<sup>+</sup> are shown as stick diagrams with green and magenta colors, respectively. (<b>a</b>) The open conformation occurs when only the NAD<sup>+</sup> cofactor is bound. The hydrogen bonding interactions between catalytic residues are presented as dotted lines. Three catalytic residues are bridged with hydrogen bonding interactions. (<b>b</b>) Upon substrate binding, the ligands and the catalytic residues adopt closed conformation. The hydrogen bonding interactions are marked as dotted lines. The distance between the NE2 of His136 and the side chain oxygen of Ser117 residue is presented with a pale-blue-colored dotted line labeled as 4.9Å. The distance became more distant from 2.7Å to 4.9Å in the transition to closed conformation, indicating breakage of hydrogen bonding. (<b>c</b>,<b>d</b>) Surface electrostatic presentation of the substrate binding site; A2HBD structures without (<b>c</b>) and with (<b>d</b>) substrate binding are presented as surface electrostatic presentations with the same orientation. The arginine 143 residue is marked as a yellow dotted circle and labeled. The black arrow indicates substrate binding pocket. The orientation of the red-colored arrow indicates the downside of the guanidinium side chain of arginine. Following substrate binding, the guanidinium side chain rotates approximately 90 degrees to align parallel to the adenine group of acetoacetyl-CoA to stabilize the substrate binding. (<b>e</b>–<b>g</b>) Tertiary crystal structure of A2HBD in complex with NAD<sup>+</sup> and acetoacetyl-CoA. (<b>e</b>) Overall tertiary complex of A2HBD. A2HBD is shown as a gray surface model. The acetoacetyl-CoA and substrate-interacting residues are shown as stick diagrams. (<b>f</b>) The acetoacetyl-CoA is presented as white and N-terminal residues and C-terminal residues interacting with the substrate. Hydrogen bonding interactions are presented as dotted lines and cation-π interaction is represented as red-colored dotted lines. (<b>g</b>) The 2fofc electron density map (1.0 σ) of acetoacetyl-CoA by SFX-data.</p>
Full article ">Figure 7
<p>(<b>a</b>) Structure-based catalytic mechanism of the acetoacetyl-CoA-catalyzed reaction. Schematic diagram of proposed catalytic mechanism in A2HBD active site. Hydride transfer and nucleophilic substitution are presented as blue-colored arrows. Polarized carbonyl carbon and oxygen are indicated as red-colored δ<sup>+</sup> and δ<sup>−,</sup>, respectively. (<b>b</b>) Binding affinity analysis of NAD<sup>+</sup>, acetoacetyl-CoA, and NAD<sup>+</sup>/acetoacetyl-CoA by MST experiments. The K<sub>d</sub> values were showed BHBD with acetoacetyl-CoA in red color; BHBD in NAD<sup>+</sup>, and acetoacetyl-CoA in cyan color; BHBD with NAD<sup>+</sup> in yellow color, respectively. Each K<sub>d</sub> value is shown in the table.</p>
Full article ">
13 pages, 2386 KiB  
Article
Tsg101 UEV Interaction with Nedd4 HECT Relieves E3 Ligase Auto-Inhibition, Promoting HIV-1 Assembly and CA-SP1 Maturation Cleavage
by Susan M. Watanabe, David A. Nyenhuis, Mahfuz Khan, Lorna S. Ehrlich, Irene Ischenko, Michael D. Powell, Nico Tjandra and Carol A. Carter
Viruses 2024, 16(10), 1566; https://doi.org/10.3390/v16101566 - 2 Oct 2024
Viewed by 435
Abstract
Tsg101, a component of the endosomal sorting complex required for transport (ESCRT), is responsible for recognition of events requiring the machinery, as signaled by cargo tagging with ubiquitin (Ub), and for recruitment of downstream acting subunits to the site. Although much is known [...] Read more.
Tsg101, a component of the endosomal sorting complex required for transport (ESCRT), is responsible for recognition of events requiring the machinery, as signaled by cargo tagging with ubiquitin (Ub), and for recruitment of downstream acting subunits to the site. Although much is known about the latter function, little is known about its role in the earlier event. The N-terminal domain of Tsg101 is a structural homologue of Ub conjugases (E2 enzymes) and the protein associates with Ub ligases (E3 enzymes) that regulate several cellular processes including virus budding. A pocket in the domain recognizes a motif, PT/SAP, that permits its recruitment. PT/SAP disruption makes budding dependent on Nedd4L E3 ligases. Using HIV-1 encoding a PT/SAP mutation that makes budding Nedd4L-dependent, we identified as critical for rescue the residues in the catalytic (HECT) domain of the E3 enzyme that lie in proximity to sites in Tsg101 that bind Ub non-covalently. Mutation of these residues impaired rescue by Nedd4L but the same mutations had no apparent effect in the context of a Nedd4 isomer, Nedd4-2s, whose N-terminal (C2) domain is naturally truncated, precluding C2-HECT auto-inhibition. Surprisingly, like small molecules that disrupt Tsg101 Ub-binding, small molecules that interfered with Nedd4 substrate recognition arrested budding at an early stage, supporting the conclusion that Tsg101–Ub–Nedd4 interaction promotes enzyme activation and regulates Nedd4 signaling for viral egress. Tsg101 regulation of E3 ligases may underlie its broad ability to function as an effector in various cellular activities, including viral particle assembly and budding. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">(</span><b>A</b>) Frequencies of contacts observed between the Tsg101 UEV domain and the HECT domain of Nedd4. Contact frequencies are derived from the top 10% of ensemble pairs taken from modeling of solution NMR paramagnetic relaxation enhancement experiments [<a href="#B20-viruses-16-01566" class="html-bibr">20</a>] for labels at sites 528 and 867 in the Nedd4 HECT domain (PDB ID: 57CJ, ”L” orientation shown). Contacts were defined as a residue of Tsg101 UEV coming within 4 Angstroms of the corresponding HECT residue, where HECT residue numbers are given both for Nedd4 (top) and the equivalent position in Nedd4L (bottom). Two regions with the most contacts; αH1 and the region proximal to the catalytic cysteine in the C-lobe are highlighted. As Tsg101 UEV is structurally homologous to an E2 enzyme, interaction in this region far from the canonical E2 binding site was unexpected. (<b>B</b>) A view of the Nedd4L HECT domain taken from PDB ID: 3JVZ highlighting the region around αH-1 (blue). The determinants in this region (Y601 and Y603) investigated here are highlighted in orange.</p>
Full article ">Figure 2
<p>Mutation of Nedd4L HECT residues in UEV proximity prevented virus rescue. (<b>A</b>) 293T cells were transfected with pNL4-LIRL alone (lanes 1,5) or co-transfected with 50–200 ng Nedd4L WT (lanes 2–4) or Nedd4L YY601,603AA mutant (lanes 6–8). Top panel, levels of VLP detected by Western blots; bottom panel, Gag p24 and p25 in cell lysates. (<b>B</b>) Western blot signals were quantified relative to the pNL4-LIRL samples with no Nedd4. Top panel, VLP produced when co-transfected with Nedd4L WT or YY601,603AA mutant; bottom panel, the ratio of Gag p24 and p25 levels in cell lysates co-transfected with Nedd4L WT or YY601,603AA. The levels of VLP produced and the ratio of Gag p24 and p25 in the cell lysate were significantly different when co-transfected with the Nedd4L Y601/Y603AA mutant versus Nedd4L WT (Student’s <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.01). Number of independent trials, <span class="html-italic">n</span> = 2.</p>
Full article ">Figure 3
<p>C2 domain truncation compensated for mutation of HECT residues in UEV proximity. (<b>A</b>) 293T cells were transfected with pNL4-LIRL alone (lanes 1,5) or co-transfected with 50–200 ng Nedd4-2s WT which has a naturally occurring C2 domain truncation relative to Nedd4L (lanes 2–4) or Nedd4-2s YY601,603AA mutant (lanes 6–8). Top panel, levels of VLP detected by Western blots; bottom panel, Gag p24 and p25 produced in cell lysates. (<b>B</b>) Western blot signals from rescue assays were quantified relative to the pNL4-LIRL samples with no Nedd4-2s assayed in parallel. Top panel, VLP produced when co-transfected with Nedd4-2s WT or YY601,603AA mutant; bottom panel, Gag p24/p25 ratios in cell lysates co-transfected with Nedd4-2s WT or YY601,603AA. The increase in VLP production and the ratio of p24/p25 for Nedd4-2s YY601,603AA were not significantly different from Nedd4-2s WT (Student’s <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &gt; 0.05). Number of independent trials, <span class="html-italic">n</span> = 2.</p>
Full article ">Figure 4
<p>Deletion of the HECT domain αH1 abrogates virus rescue. (<b>A</b>) 293T cells were transfected with pNL4-3-LIRL alone (lanes 1,5) or co-transfected with 50–200 ng Nedd4-2s WT (lanes 2–4) or Nedd4-2s-ΔαH1 (lanes 6–8). The Nedd4-2s constructs were compared for their ability to rescue pNL4-LIRL budding as measured by the levels of VLP detected by Western blots (top panel) and Gag p24 and p25 in cell lysates (bottom panel). (<b>B</b>) Western blot signals from rescue assays were quantified relative to the pNL4-LIRL samples with no Nedd4-2s assayed in parallel. Top panel, VLP produced when co-transfected with Nedd4-2s WT or Nedd4-2s-ΔαH1; bottom panel, the ratio of Gag p24 and p25 for Nedd4-2s WT or Nedd4-2s-ΔαH1. The VLP efficiency and the ratio of p24/p25 for ΔαH1 mutant were significantly different from WT (Student’s <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.01). Number of independent trials, <span class="html-italic">n</span> = 2.</p>
Full article ">Figure 5
<p>Benserazide (K21) arrests budding of pNL4-3-P7L. (<b>A</b>) WST-1 assay of cell metabolic activity at increasing concentrations of K21 (CC50 110–140 µM, 95% CI); (<b>B</b>) Effect of K21 (50 µM) on VLP production. Top panel, Western analysis of virus-like particles (VLPs); bottom panel, Gag proteins in cell lysate for 293T cells transfected with pNL4-3 WT. (<b>C</b>) Effect of early versus late K21 (50 µM) addition on VLP production as determined by ELISA and MAGI assays (<span class="html-italic">n</span> = 2). Numbers were normalized to the untreated (DMSO) controls. (<b>D</b>) Examination by electron microscopy of mock, K21-treated cells, and Tenatoprazole-treated cells transfected with NL4-3-P7L (top). Bottom, Quantitative analysis of budding morphologies in cells exposed to either DMSO or 50 µM K21. For P7L, distribution of morphologies for control (DMSO) versus K21 were significantly different (Chi square test, <span class="html-italic">p</span> &lt; 0.001); for WT, distribution was not significantly affected by K21.</p>
Full article ">Figure 6
<p>Model: Tsg101 UEV interference with Nedd4 autoinhibition. Nedd4L is normally in an autoinhibited, inactive state (top left) induced by backbinding of the N-terminal C2 domain (blue) and WW domains (orange) to the catalytic HECT domain (gray). This form exists alongside the active form (top right), which may be induced by interaction of the Tsg101 UEV domain (pink), which in turn may interact with Ub (yellow) on the HECT domain or through direct interaction with the HECT domain or αH1 region (green). Amino acid substitutions in αH1 (bottom left) presumptively alleviates the autoinhibited state and Tsg101 interaction, similarly to the deletion of the entire αH1 region (bottom right). Mutation or deletion of αH1 has no impact on the Nedd4-2s isoform (right panel), whose truncated C2 domain putatively prevents backbinding and formation of the autoinhibited state. The Nedd4-2s isoform is distinct from the FL in its ability to interact with HIV-1 Gag, whose binding potentially affects helix-1 (far right, green).</p>
Full article ">
14 pages, 2491 KiB  
Technical Note
A Bacterial Platform for Studying Ubiquitination Cascades Anchored by SCF-Type E3 Ubiquitin Ligases
by Zuo-Xian Pu, Jun-Li Wang, Yu-Yang Li, Luo-Yu Liang, Yi-Ting Tan, Ze-Hui Wang, Bao-Lin Li, Guang-Qin Guo, Li Wang and Lei Wu
Biomolecules 2024, 14(10), 1209; https://doi.org/10.3390/biom14101209 - 25 Sep 2024
Viewed by 492
Abstract
Ubiquitination is one of the most important post-translational modifications in eukaryotes. The ubiquitination cascade includes ubiquitin-activating enzymes (E1), ubiquitin-conjugating enzymes (E2), and ubiquitin ligases (E3). The E3 ligases, responsible for substrate recognition, are the most abundant and varied proteins in the cascade and [...] Read more.
Ubiquitination is one of the most important post-translational modifications in eukaryotes. The ubiquitination cascade includes ubiquitin-activating enzymes (E1), ubiquitin-conjugating enzymes (E2), and ubiquitin ligases (E3). The E3 ligases, responsible for substrate recognition, are the most abundant and varied proteins in the cascade and the most studied. SKP1-CUL1-F-Box (SCF)-type E3 ubiquitin ligases are multi-subunit RING (Really Interesting New Gene) E3 ubiquitin ligases, composed of CUL1 (Cullin 1), RBX1 (RING BOX 1), SKP1 (S-phase Kinase-associated Protein 1), and F-box proteins. In vitro ubiquitination assays, used for studying the specific recognition of substrate proteins by E3 ubiquitin ligases, require the purification of all components involved in the cascade, and for assays with SCF-type E3 ligases, additional proteins (several SCF complex subunits). Here, the Duet expression system was used to co-express E1, E2, ubiquitin, ubiquitylation target (substrate), and the four subunits of a SCF-type E3 ligase in E. coli. When these proteins co-exist in bacterial cells, ubiquitination occurs and can be detected by Western Blot. The effectiveness of this bacterial system for detecting ubiquitination cascade activity was demonstrated by replicating both AtSCFTIR1-mediated and human SCFFBXO28-mediated ubiquitylation in bacteria. This system provides a basic but adaptable platform for the study of SCF-type E3 ubiquitin ligases. Full article
(This article belongs to the Section Biomacromolecules: Proteins)
Show Figures

Figure 1

Figure 1
<p>Expression of all components of the ubiquitination cascade by the <span class="html-italic">Duet</span> vector system. (<b>A</b>) The structures of <span class="html-italic">pCDFDuet-AtUBC8-S-AtUBA1-S</span>, <span class="html-italic">pRSFDuet-His-Flag-AtUBQ11</span>, and <span class="html-italic">pETDuet-AtRBX1-T7-AtASK1-T7-AtCUL1-S</span>. AtUBC8 and AtUBA1 are S-tagged at their C-termini. AtUBQ11 is His-FLAG-tagged at its N-terminus. AtRBX1 and AtASK1 are T7-tagged at their C-termini. AtCUL1 is S-tagged at its C-terminus. (<b>B</b>–<b>E</b>) The <span class="html-italic">E. coli</span> strains BL21(DE3) containing the individual plasmids in A were cultured separately. Crude proteins obtained by lysing these bacteria via ultrasonication were used to detect recombinant proteins using immunoblot. (<b>F</b>) The vector <span class="html-italic">pACYCDuet-Myc-MBP-HA</span> designed for the expression of the Myc-tagged F-box protein and MBP-HA-tagged substrate protein. (<b>G</b>) After obtaining crude proteins from the lysed bacteria containing the plasmid in F, an HA-blot is used to detect MBP-HA. In B-E and G, crude proteins from strains not induced by IPTG were used for negative controls, with some recombinant proteins showing a weak, leaky expression. (<b>A</b>,<b>F</b>) Turquoise represents CDS of different genes, orange represents S-tagged, blue represents His-tagged, purple represents Flag-tagged, pink represents T7-tagged, yellow represents Myc-tagged, cyan represents MBP-tagged, and light green represents HA-tagged. Original images can be found in <a href="#app1-biomolecules-14-01209" class="html-app">Figure S1</a>.</p>
Full article ">Figure 2
<p>Reconstitution of AtSCF<sup>TIR1</sup> auto-ubiquitination in <span class="html-italic">Escherichia coli</span> bacteria. (<b>A</b>) The structures of <span class="html-italic">pACYCDuet-AtTIR1-Myc-MBP-HA</span> and <span class="html-italic">pACYCDuet-AtTIR1<sup>P10A</sup>-Myc-MBP-HA</span>. AtTIR1 or AtTIR1<sup>P10A</sup> are Myc-tagged at their C-termini. (<b>B</b>) The structures of <span class="html-italic">pCDFDuet-AtUBA1-S</span> and <span class="html-italic">pCDFDuet-AtUBC8-S</span>. AtUBA1 and AtUBC8 are S-tagged at their C-termini. (<b>C</b>) The auto-ubiquitination of AtSCF<sup>TIR1</sup> is detected after the co-expression of all components of the ubiquitination cascade, except the substrate in <span class="html-italic">E coli</span>. The strains missing one or more of these components served as the negative controls. The mutation of AtTIR1<sup>P10A</sup>-Myc co-expressed with other components of the ubiquitination cascade served as another negative control. Auto-ubiquitination activities of AtSCF<sup>TIR1</sup> were analyzed by Western Blot with anti-Myc or anti-ubiquitin antibodies. AtUBA1, AtUBC8, and AtCUL1 were analyzed by WB with anti-S antibodies. AtRBX1 and AtASK1 were analyzed by WB with anti-T7 antibodies. (<b>A</b>,<b>B</b>) Turquoise represents CDS of different genes, orange represents S-tagged, yellow represents Myc-tagged, cyan represents MBP-tagged, and light green represents HA-tagged. Original images can be found in <a href="#app1-biomolecules-14-01209" class="html-app">Figure S2</a>.</p>
Full article ">Figure 3
<p>Ubiquitination of AtIAA6 by AtSCF<sup>TIR1</sup> in <span class="html-italic">Escherichia coli</span> bacteria. (<b>A</b>) The structures of <span class="html-italic">pACYCDuet-AtTIR1-Myc-MBP-AtIAA6-HA</span> and <span class="html-italic">pACYCDuet-Myc-MBP-AtIAA6-HA</span>. AtTIR1 is Myc-tagged at its C-terminus. AtIAA6 is MBP-tagged at its N-terminus and HA-tagged at its C-terminus. Turquoise represents CDS of different genes, orange represents S-tagged, yellow represents Myc-tagged, cyan represents MBP-tagged, and light green represents HA-tagged. (<b>B</b>) The ubiquitination of AtIAA6 is detected after the co-expression of all components of the ubiquitination cascade in <span class="html-italic">E coli</span>. The strains missing one or more of these components served as negative controls. Auxin IAA was added to all expression systems, except to the control. The ubiquitination of AtIAA6 was analyzed by WB with anti-HA. The auto-ubiquitination of AtTIR1 was analyzed by WB with anti-Myc. The ubiquitination of AtIAA6 and AtTIR1 was detected simultaneously using anti-ubiquitin. AtUBA1, AtUBC8, and AtCUL1 were analyzed by WB with anti-S. AtRBX1 and AtASK1 were analyzed by WB with anti-T7. Original images can be found in <a href="#app1-biomolecules-14-01209" class="html-app">Figure S3</a>.</p>
Full article ">Figure 4
<p>Reconstitution of SCF<sup>FBXO28</sup> auto-ubiquitination in <span class="html-italic">Escherichia coli.</span> (<b>A</b>) The structures of <span class="html-italic">pETDuet-AtRBX1-T7-HsSKP1-T7-AtCUL1-S, pACYCDuet-HsFBXO28-Myc-MBP-HA</span>, and <span class="html-italic">pACYCDuet-ΔHsFBXO28-Myc-MBP-HA</span>. The human protein HsHFBXO28 and its inactive form ΔHsFBXO28 are Myc-tagged at their C-termini. Turquoise represents CDS of different genes, orange represents S-tagged, purple represents Flag-tagged, pink represents T7-tagged, yellow represents Myc-tagged, cyan represents MBP-tagged, and light green represents HA-tagged. (<b>B</b>) The auto-ubiquitination of SCF<sup>FBXO28</sup> is detected after the co-expression of all components of the ubiquitination cascade, except substrate in <span class="html-italic">E coli</span>. SCF<sup>FBXO28</sup> was composed of heterologous subunits: AtCUL1, AtRBX1, HsSKP1, and HsFBXO28. The strains missing one or more of these components served as negative controls. Mutated ΔHsFBXO28-Myc served as another negative control. Auto-ubiquitination activities of SCF<sup>FBXO28</sup> were analyzed by WB with anti-Myc or anti-ubiquitin. AtUBA1, AtUBC8, and AtCUL1 were analyzed by WB with anti-S. AtRBX1 and HsSKP1 were analyzed by WB with anti-T7. Original images can be found in <a href="#app1-biomolecules-14-01209" class="html-app">Figure S4</a>.</p>
Full article ">
19 pages, 4525 KiB  
Article
Mechanistic Insights into Substrate Recognition of Human Nucleoside Diphosphate Kinase C Based on Nucleotide-Induced Structural Changes
by Rezan Amjadi, Sebastiaan Werten, Santosh Kumar Lomada, Clara Baldin, Klaus Scheffzek, Theresia Dunzendorfer-Matt and Thomas Wieland
Int. J. Mol. Sci. 2024, 25(18), 9768; https://doi.org/10.3390/ijms25189768 - 10 Sep 2024
Viewed by 465
Abstract
Nucleoside diphosphate kinases (NDPKs) are encoded by nme genes and exist in various isoforms. Based on interactions with other proteins, they are involved in signal transduction, development and pathological processes such as tumorigenesis, metastasis and heart failure. In this study, we report a [...] Read more.
Nucleoside diphosphate kinases (NDPKs) are encoded by nme genes and exist in various isoforms. Based on interactions with other proteins, they are involved in signal transduction, development and pathological processes such as tumorigenesis, metastasis and heart failure. In this study, we report a 1.25 Å resolution structure of human homohexameric NDPK-C bound to ADP and describe the yet unknown complexes formed with GDP, UDP and cAMP, all obtained at a high resolution via X-ray crystallography. Each nucleotide represents a distinct group of mono- or diphosphate purine or pyrimidine bases. We analyzed different NDPK-C nucleotide complexes in the presence and absence of Mg2+ and explain how this ion plays an essential role in NDPKs’ phosphotransferase activity. By analyzing a nucleotide-depleted NDPK-C structure, we detected conformational changes upon substrate binding and identify flexible regions in the substrate binding site. A comparison of NDPK-C with other human isoforms revealed a strong similarity in the overall composition with regard to the 3D structure, but significant differences in the charge and hydrophobicity of the isoforms’ surfaces. This may play a role in isoform-specific NDPK interactions with ligands and/or important complex partners like other NDPK isoforms, as well as monomeric and heterotrimeric G proteins. Considering the recently discovered role of NDPK-C in different pathologies, these high-resolution structures thus might provide a basis for interaction studies with other proteins or small ligands, like activators or inhibitors. Full article
Show Figures

Figure 1

Figure 1
<p>Overall view of human NDPK-C in complex with ADP—The homo-hexameric structure of NDPK-C can be considered as a trimer of dimers. (<b>A</b>) Top view—the mainly visible subunits oriented with threefold rotational symmetry are colored in blue, green and orange, respectively; the nucleotide ADP is shown in light blue. (<b>B</b>) Side view—generated by a 90 degree rotation around the horizontal axis, with same colors as in A. Here, the corresponding dimer-forming protomers are visible and colored in light blue, light green and light orange, respectively. (<b>C</b>) Top view—with all assigned secondary structure elements indicated and labeled in protomer 1 (blue) as in A. Only elements involved in the trimer contacts between the basic dimers are labeled in protomers 2 and 3 (green and orange). (<b>D</b>) Side view (orientation as in B) indicating the elements forming the dimer interface. (<b>E</b>) Zoomed-in view of the dimer interface, which is formed by an augmentation via β3 strands and additionally stabilized by salt bridges and hydrogen bonds (shown in dashed lines). (<b>F</b>) Analysis of the NDPK-C oligomeric state in solution. Chromatogram of an analytical native gel filtration analysis with refractive index (RI) and multiple-angle light scattering (MALS) detection. The calculated molar masses (left axis) are shown as dotted lines for NDPK-C (blue) and a control protein (Bovine serum albumin, shown in grey). Masses corresponding to concentration peak maxima (RI based solid lines, right axis) are displayed.</p>
Full article ">Figure 2
<p>Human NDPK-C×ADP structure in the absence and presence of Mg<sup>2+</sup>—(<b>A</b>) the ribbon model (blue) shows protomer 1 of the magnesium-free hexameric NDPK-C structure in complex with ADP (space group <span class="html-italic">P</span>1). The Fo-Fc electron density map of the substrate contoured at σ = 1.3 is shown as a gray mesh together with the modeled ADP in atomic coloring. The nucleotide-binding site (boxed) is shown in a close-up view below. Lys29, Phe77, Arg105, Arg122, Val129 and Asn132 form the ADP-binding pocket (3 Å radius). (<b>B</b>) Stabilization of ADP in the active site of NDPK-C via indirect interaction with three water molecules (red spheres) bridging Tyr69, His135 and Gly136. (<b>C</b>) The ribbon model (pink) shows one NDPK-C subunit in the presence of magnesium (space group <span class="html-italic">C</span>2). The corresponding Fo-Fc map contoured at σ = 1.2 includes additional density that could be assigned to a magnesium ion and three water molecules (green and red spheres, respectively). The distance of Arg122 is longer than that in A (4 Å). (<b>D</b>) ADP stabilization in the presence of Mg<sup>2+</sup> (green sphere) also includes Asp138, which is close to the β-phosphate and one of the four water molecules (red spheres). (<b>E</b>) Superposition of the two nucleotide pockets (A and C) show a change in the conformation of the phosphate chain of ADP upon metal ion binding. The black arrow indicates the movement of the β-phosphate toward the Mg<sup>2+</sup> ion (with a respective distance of 1 Å) and the weakening of its interaction with Arg122. (<b>F</b>) Superposition of human NDPK-C (pink-and-green Mg<sup>2+</sup> and four water molecules in red) with a <span class="html-italic">D. discoideum</span> NDPK-ADP-AlF<sub>3</sub> complex (1KDN in grey with light-green Mg<sup>2+</sup> and light-brown water molecules) showing the metal ion in octahedral coordination.</p>
Full article ">Figure 3
<p>NDPK-C binding mode for different nucleotides (<b>A</b>) Two hexamers of NDPK-C in complex with GDP are linked via a bridging SO<sub>4</sub><sup>2−</sup> between subunits from each hexamer. The magnification view (below) shows the interface of two subunits. We observed contacts between Arg80 from each hexamer and Pro76 (cyan) with Asn112 from the other hexamer (light cyan). (<b>B</b>) Electron density map for the ligand GDP and superposition of NDPK-C×ADP (blue) with NDPK-C×GDP (cyan) complexes. In the GDP complex, the C2 amino group of the guanine base interacts with Glu169 from the neighboring chain inside the hexamer. (<b>C</b>) Electron density map for the ligand UDP and superposition of NDPK-C×UDP on NDPK-C×ADP (space group <span class="html-italic">C</span>2 with magnesium) shows that Phe77 shows a similar π-stack with purine and pyrimidine bases. (<b>D</b>) Electron density map for the ligand cAMP and active site of human NDPK-C in complex with cAMP. Lys29, Tyr69, Phe77, Val129, Asn132 and His135 directly interact with cAMP, while Arg105, Arg122 and Gly136 are involved in interaction with cAMP via a water molecule (shown as red sphere). (<b>E</b>) Superposition of NDPK-C×cAMP with NDPK-C×ADP complexes shows the different orientation and interactions of cAMP (pink) with active site residues compared to ADP (blue). In contrast to the ADP coordination the distance between the α-phosphate of cAMP and the Kpn-loop is larger, and this phosphate is closer to Tyr69 and the catalytically active His135. (<b>F</b>) NDPK-C activity in vitro. Upper panel: Enzymatic activity of recombinant NDPK-C (2 nM) was analyzed in the absence (spheres) or presence of cAMP at 1 mM (squares) or 5 mM (triangles); samples were taken at the indicated time points. Formation of ATP from ADP and GTP was analyzed by quantitative ion-pair chromatography (HPLC). Lower panel: Similar experiments were performed at low Mg<sup>2+</sup> concentration (20 nM, spheres) or in the presence of 5 mM EDTA (squares); samples were taken at the indicated time points. Formation of GDP and ATP was analyzed by quantitative ion-pair chromatography (HPLC); data are shown for ATP accumulation. Right: Purified NDPK-C visualized with SYPRO Ruby Protein Gel Stain (upper panel) and the autophosphorylation of the catalytic histidine residue in the presence of Mg<sup>2+</sup>-ATP as detected with the anti-1p-His-antibody (lower panel).</p>
Full article ">Figure 4
<p>Structure of nucleotide-depleted human NDPK-C (<b>A</b>) Ribbon model of one nucleotide-depleted NDPK-C subunit (green); only protomer 1 of the hexameric structure is shown for clarity. The Fo-Fc electron density map of the phosphate ion and water molecule contoured at σ = 1.3 is shown as a grey mesh together with the modeled HPO<sub>4</sub><sup>2−</sup> in orange sticks. The phosphate is stabilized by interaction with Lys29, Tyr69, Asn132, His135 and a water molecule that is linked to Arg105 and Gly136. (<b>B</b>) Superposition of nucleotide-depleted (green) with one NDPK-C×UDP subunit (yellow) shows the movement of helix α2 (including Ala60 to Arg73) in the presence and absence of UDP. The distance between Val129 (located in Kpn-loop) and Phe77 (located in helix α2), which can be considered as a gate to access the active site, is increased from 10.2 Å in the NDPK-C×UDP complex to 11.3 Å in depleted-UDP NDPK-C. An enlarged view of NDPK-C active site in UDP (yellow) and nucleotide-free structures (green) shows that the free phosphate ion is localized in the same area as α-phosphate in NDPK-C×UDP structure but closer to His135 to establish direct interaction. (<b>C</b>) The value of the B-factor in the UDP-depleted structure is increased in both helices (αA and α2) compared with NDPK-C in complex with UDP. Increasing the B-factor in this area resembles the movement of the respective region in depleted-nucleotide NDPK-C (green plot). (<b>D</b>) Superposition of nucleotide-depleted NDPK-C (green) and NDPK from <span class="html-italic">D. discoideum</span> (grey 1KDN). The bound AlF<sub>3</sub> forms an accurate analogue of the transition of γ-phosphate in NDPK×ADP×AlF<sub>3</sub> complex from <span class="html-italic">D. discoideum</span> (grey). The superposition of nucleotide-depleted NDPK-C (green) and NDPK from <span class="html-italic">D. discoideum</span> (grey 1KDN) shows the phosphate ion in depleted NDPK-C is positioned in a very similar place as AlF3 in NDPK from <span class="html-italic">D. discoideum</span>, which can be viewed as a transition state of NDPK-C. Water molecules are shown as red spheres. (<b>E</b>) Superposition of depleted NDPK-C (green) and NDPK from <span class="html-italic">D. discoideum</span> (grey 1NSP) shows the similar localization of phosphate in depleted-nucleotide NDPK-C as phosphate, which is covalently bound to the catalytic histidine in <span class="html-italic">D. discoideum</span> NDPK.</p>
Full article ">Figure 5
<p>NDPK-B and -C isoforms show differences in surface potential—(<b>A</b>) Primary sequence alignment of NDPK-B and -C orthologues from the indicated species. Conserved residues are shaded in gray. Residues which differ between the isoforms but are specifically conserved within one isoform are highlighted in yellow and brown, respectively. Residue positions that (i) show isoform-specific conservation; (ii) are located on the surface of the hexameric 3D structure; and (iii) are chemically different in terms of their isoforms are indicated with stars: Gly63/Arg80, Asp148/His165 and Lys143/Glu160. (<b>B</b>) Surface display of NDPK-B (PDB 1NUE) with conserved isoform-specific residues labeled using atomic coloring. (<b>C</b>) Surface view of NDPK-C with conserved residues of Arg80, His165 and Glu160 highlighted on the surface. The two isoforms show clear differences in the location of charged residues on their surface.</p>
Full article ">
19 pages, 8433 KiB  
Article
Validation of In-House Imaging System via Code Verification on Petunia Images Collected at Increasing Fertilizer Rates and pHs
by Kahlin Wacker, Changhyeon Kim, Marc W. van Iersel, Mark Haidekker, Lynne Seymour and Rhuanito Soranz Ferrarezi
Sensors 2024, 24(17), 5809; https://doi.org/10.3390/s24175809 - 6 Sep 2024
Viewed by 625
Abstract
In a production environment, delayed stress recognition can impact yield. Imaging can rapidly and effectively quantify stress symptoms using indexes such as normalized difference vegetation index (NDVI). Commercial systems are effective but cannot be easily customized for specific applications, particularly post-processing. We developed [...] Read more.
In a production environment, delayed stress recognition can impact yield. Imaging can rapidly and effectively quantify stress symptoms using indexes such as normalized difference vegetation index (NDVI). Commercial systems are effective but cannot be easily customized for specific applications, particularly post-processing. We developed a low-cost customizable imaging system and validated the code to analyze images. Our objective was to verify the image analysis code and custom system could successfully quantify the changes in plant canopy reflectance. ‘Supercascade Red’, ‘Wave© Purple’, and ‘Carpet Blue’ Petunias (Petunia × hybridia) were transplanted individually and subjected to increasing fertilizer treatments and increasing substrate pH in a greenhouse. Treatments for the first trial were the addition of a controlled release fertilizer at six different rates (0, 0.5, 1, 2, 4, and 8 g/pot), and for the second trial, fertilizer solution with four pHs (4, 5.5, 7, and 8.5), with eight replications with one plant each. Plants were imaged twice a week using a commercial imaging system for fertilizer and thrice a week with the custom system for pH. The collected images were analyzed using an in-house program that calculated the indices for each pixel of the plant area. All cultivars showed a significant effect of fertilizer on the projected canopy size and dry weight of the above-substrate biomass and the fertilizer rate treatments (p < 0.01). Plant tissue nitrogen concentration as a function of the applied fertilizer rate showed a significant positive response for all three cultivars (p < 0.001). We verified that the image analysis code successfully quantified the changes in plant canopy reflectance as induced by increasing fertilizer application rate. There was no relationship between the pH and NDVI values for the cultivars tested (p > 0.05). Manganese and phosphorus had no significance with chlorophyll fluorescence for ‘Carpet Blue’ and ‘Wave© Purple’ (p > 0.05), though ‘Supercascade Red’ was found to have significance (p < 0.01). pH did not affect plant canopy size. Chlorophyll fluorescence pixel intensity against the projected canopy size had no significance except in ‘Wave© Purple’ (p = 0.005). NDVI as a function of the projected canopy size had no statistical significance. We verified the ability of the imaging system with integrated analysis to quantify nutrient deficiency-induced variability in plant canopies by increasing pH levels. Full article
(This article belongs to the Section Smart Agriculture)
Show Figures

Figure 1

Figure 1
<p>Flowchart diagram of the in-house imaging system to capture and analyze plant images under different light-emitting diodes (LEDs) wavelengths using chlorophyll fluorescence imaging to calculate spatial NDVI and canopy size per pixel for detailed plant analysis.</p>
Full article ">Figure 2
<p>Details of each image obtained by the imaging system, histogram representation, and normalized difference vegetation (NDVI) and anthocyanin content index (ACI) false color images.</p>
Full article ">Figure 3
<p>Projected canopy size of three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown under increasing fertilizer rates. The fertilizer rate applied has a significant effect on the two-dimensional area of the plant, as measured by a commercial imaging system and analyzed by our in-house software. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 4
<p>Dry mass of three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown under increasing fertilizer rates. All cultivars show significance in the treatments. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 5
<p>Nitrogen concentration as a function of increasing fertilizer rate on three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia). The nitrogen concentration was shown to be significantly related to the fertilizer rate. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 6
<p>Normalized difference vegetation index (NDVI) from the imaging system for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) at increasing fertilizer application rates. The normalized difference vegetation index (NDVI) responses are shown to be significantly related to fertilizer application. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 7
<p>Plant tissue nitrogen concentration as a function of the average pixel normalized difference vegetation index (NDVI) of the plant area for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) subjected to increasing fertilizer rates.</p>
Full article ">Figure 8
<p>Projected canopy size in pixels against the tissue nitrogen concentration for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown at increasing fertilizer rates. Primarily, this shows the effect of nitrogen concentration on the plant growth size.</p>
Full article ">Figure 9
<p>Dry biomass as a function of the projected canopy size for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown at increasing fertilizer rates. This shows the correlation between the imaged plant size and the dry mass.</p>
Full article ">Figure 10
<p>Projected canopy as a function of normalized difference vegetation index (NDVI), both from imaging system for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown at increasing fertilizer rates.</p>
Full article ">Figure 11
<p>(<b>A</b>) Phosphorus and (<b>B</b>) Manganese concentrations of ‘Supercascade Red’ in response to increasing pH. These nutrient decreases in the plant tissue were the desired effect in the experiment to display deficiencies or other visible symptoms. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 12
<p>Normalized difference vegetation index (NDVI) response to pH for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Normalized difference vegetation index (NDVI) did not show a meaningful response to pH, except for ‘Supercascade Red’, which could be considered significant due to several extreme outliers. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 13
<p>Manganese content against chlorophyll fluorescence for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. There was no significant effect of Manganese on image-measured parameters on ‘Carpet Blue’ and ‘Wave© Purple’ cultivars.</p>
Full article ">Figure 14
<p>Phosphorus concentration against chlorophyll fluorescence for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Stronger effect with phosphorus, explained by phosphorus being a macronutrient rather than a micronutrient.</p>
Full article ">Figure 15
<p>Average chlorophyll fluorescence pixel intensity as a function of pH for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 16
<p>Projected canopy size as a function of the pH treatments for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions. Each point is the mean of 8 replicates with standard error bars.</p>
Full article ">Figure 17
<p>Chlorophyll fluorescence as a function of the projected canopy size for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions.</p>
Full article ">Figure 18
<p>Normalized difference vegetation index (NDVI) plotted against the projected canopy size for three cultivars of petunia (<span class="html-italic">Solanaceae Petunia</span> × hybridia) grown in increasing pH solutions.</p>
Full article ">
17 pages, 9841 KiB  
Article
Elucidating the Substrate Envelope of Enterovirus 68-3C Protease: Structural Basis of Specificity and Potential Resistance
by Vincent N. Azzolino, Ala M. Shaqra, Akbar Ali, Nese Kurt Yilmaz and Celia A. Schiffer
Viruses 2024, 16(9), 1419; https://doi.org/10.3390/v16091419 - 5 Sep 2024
Viewed by 587
Abstract
Enterovirus-D68 (EV68) has emerged as a global health concern over the last decade with severe symptomatic infections resulting in long-lasting neurological deficits and death. Unfortunately, there are currently no FDA-approved antiviral drugs for EV68 or any other non-polio enterovirus. One particularly attractive class [...] Read more.
Enterovirus-D68 (EV68) has emerged as a global health concern over the last decade with severe symptomatic infections resulting in long-lasting neurological deficits and death. Unfortunately, there are currently no FDA-approved antiviral drugs for EV68 or any other non-polio enterovirus. One particularly attractive class of potential drugs are small molecules inhibitors, which can target the conserved active site of EV68-3C protease. For other viral proteases, we have demonstrated that the emergence of drug resistance can be minimized by designing inhibitors that leverage the evolutionary constraints of substrate specificity. However, the structural characterization of EV68-3C protease bound to its substrates has been lacking. Here, we have determined the substrate specificity of EV68-3C protease through molecular modeling, molecular dynamics (MD) simulations, and co-crystal structures. Molecular models enabled us to successfully characterize the conserved hydrogen-bond networks between EV68-3C protease and the peptides corresponding to the viral cleavage sites. In addition, co-crystal structures we determined have revealed substrate-induced conformational changes of the protease which involved new interactions, primarily surrounding the S1 pocket. We calculated the substrate envelope, the three-dimensional consensus volume occupied by the substrates within the active site. With the elucidation of the EV68-3C protease substrate envelope, we evaluated how 3C protease inhibitors, AG7088 and SG-85, fit within the active site to predict potential resistance mutations. Full article
Show Figures

Figure 1

Figure 1
<p>EV68-3C protease substrates and structure. (<b>A</b>) EV68 polyprotein prior to cleavage by the 3C protease into structural and non-structural viral proteins. (<b>B</b>) Crystal structure of EV68-3C protease in surface representation bound to a peptidomimetic inhibitor (SG-85, in magenta sticks) with cysteine protease’s catalytic dyad residues (H40 and C147) highlighted in yellow (PDB ID: 3ZVF). The inhibitor is labeled with corresponding P5 to P1′ moieties. (<b>C</b>) Amino acid sequences of EV68-3C protease cleavage sites in the viral polyprotein, with a conserved glutamine (Q) in the P1 position, generally followed by a glycine (G) at P1′. The sequences are highly diverse distal to the cut site (denoted by the blue vertical bar).</p>
Full article ">Figure 2
<p>Hydrogen bond network between the substrate peptides and the active site of EV68-3C protease. (<b>A</b>) The EV68-3C protease in grey surface representation, focusing on the active site pocket with the catalytic dyad highlighted in yellow, and the end point modeled 3C3D peptide (8 mer) post MD simulation shown in blue sticks. Hydrogen bonds between the protein and peptide heavy atoms are depicted with green dashed lines. (<b>B</b>) Percent frequency of the hydrogen bonds between EV68-3C protease and modeled substrate peptides, as well as the 3B3C peptide co-crystal structure (PDB ID: 9AX9), during MD simulation trajectories. The bonds present during the simulation with higher frequency interactions are denoted in darker red and lower frequency interactions are in white.</p>
Full article ">Figure 3
<p>The substrate envelope of EV68-3C protease. A heatmap representation of the calculated dynamic substrate envelope for EV68-3C protease shows the percent occupancy of viral substrates within this three-dimensional volume. The peptide backbone is most well conserved between the P4 and P2′ positions and has more variability at P5 and at P3′. The percent occupancy ranges from lowest (dark blue) to highest (red) of finding atoms at that position within a given viral peptide.</p>
Full article ">Figure 4
<p>The experimental crystal structure of EV68-3C protease with a substrate peptide bound at the active site. (<b>A</b>) The EV68-3C protease (grey surface) co-crystal structure with the 3B3C peptide (purple sticks) in the electron density, with nearby crystallographic waters (red spheres) and hydrogen bond interactions (green dashed lines) with the surrounding protease residues (dark grey sticks). The mesh depicts the electron density (2F0-FC map) for the 3B3C peptide at the active site. (<b>B</b>) The EV68-3C protease (grey surface) and 3B3C peptide (purple sticks) co-crystal structure shown with the superimposed substrate envelope denoted by the cyan volume. The peptide, derived from an experimental protein crystal structure, fits completely within the calculated envelope from molecular modeling and MD simulations.</p>
Full article ">Figure 5
<p>Comparison of peptide-bound and apo EV68-3V protease crystal structures. Surface representation of (<b>A</b>) 3B3C-peptide-bound (PDB ID: 9AX9) and (<b>B</b>) apo (PDB ID: 8FL5) protease colored according to variation of backbone, assessed by the average difference in distance between the C-alpha atoms of residues (in Å) between the two structures. (<b>C</b>) Superposition of the two structures in stick representation with red arrows pointing to the largest changes in key residues between the apo (cyan) and the substrate-bound (pink) co-crystal structures. The 3B3C peptide is depicted as purple sticks.</p>
Full article ">Figure 6
<p>(<b>A</b>) Rupintrivir (AG7088), in slate blue, covalently bound within the active site of EV68-3C protease in grey (PDB: 7L8H). Hydrogen bonds between rupintrivir and the protein are highlighted as yellow dashed lines. The substrate envelope is shown in light blue (<b>B</b>) SG-85, in orange, covalently bound within the active site of EV68-3C protease (PDB: 3ZVF).</p>
Full article ">
20 pages, 5693 KiB  
Review
Model-Based Adaptive Control of Bioreactors—A Brief Review
by Velislava Lyubenova, Maya Ignatova, Dafina Zoteva and Olympia Roeva
Mathematics 2024, 12(14), 2205; https://doi.org/10.3390/math12142205 - 13 Jul 2024
Viewed by 704
Abstract
This article summarizes the authors’ experiences in the development and application of the General Dynamical Model Approach related to adaptive linearizing control of biotechnological processes. Special attention has been given to some original, innovative solutions in model-based process control theory: new formalization of [...] Read more.
This article summarizes the authors’ experiences in the development and application of the General Dynamical Model Approach related to adaptive linearizing control of biotechnological processes. Special attention has been given to some original, innovative solutions in model-based process control theory: new formalization of biotechnological process kinetics, derivation and tuning of the general software sensor of the full kinetics of biotechnological processes, and a general algorithm for fully adaptive linearizing control with software sensors. These theoretical solutions are the basis of three control strategies—fully adaptive control of the main substrate, partially adaptive control of intermediate metabolite, and recognition and stabilization of the desired physiological state based on the proposed theoretical solutions. Each strategy is illustrated in different case studies. The advantages and limitations of each of them are identified and discussed. The derived algorithms for monitoring and controlling the considered biotechnological processes are realized and included in a software platform named Interactive System for Education in Modelling and Control of Bioprocesses (InSEMCoBio). The InSEMCoBio modules and their main functions are discussed. The effectiveness of the proposed control strategies (achieving maximum productivity) has been proven through a series of simulation investigations of the considered case studies. Full article
Show Figures

Figure 1

Figure 1
<p>Modes of cultivation of bioprocesses: <span class="html-italic">X</span>, <span class="html-italic">S</span>, <span class="html-italic">P</span>—biomass, substrate, and product concentrations [g/l], respectively; <span class="html-italic">F</span>—substrate feed rate [g/lh]; <span class="html-italic">S<sub>in</sub></span>—concentration of the substrate in the feed rate [g/l]; <span class="html-italic">V</span>—bioreactor volume [l].</p>
Full article ">Figure 2
<p>Scheme of linearizing control.</p>
Full article ">Figure 3
<p>Classic GDM approach and the proposed developments—a comparison.</p>
Full article ">Figure 4
<p>Simulation investigation of the control algorithms (15) and (16): experimental data: o glucose, + gluconic acid, * biomass, all controlled variables, and <span class="html-italic">D</span> (red line).</p>
Full article ">Figure 5
<p>Simulation investigation of the control algorithm (20): (<b>a</b>) Model and estimated values of lactate production (positive values) and consumption (negative values) rates; (<b>b</b>) Values of lactate concentration; (<b>c</b>) Difference between the estimated values of production and consumption rates; (<b>d</b>) Values of the controlled variable.</p>
Full article ">Figure 6
<p>Process dynamics in batch phase: (<b>a</b>) comparison between the glucose production rate and the consumption rate, (<b>b</b>) glucose concentration in the culture broth, and (<b>c</b>) concentration of the produced ethanol.</p>
Full article ">Figure 7
<p>Ethanol concentration (<b>left</b>) and ethanol production rate (<b>right</b>): controlled fed-batch vs. batch process.</p>
Full article ">Figure 8
<p>Scheme of software sensors designed for monitoring of the physiological states.</p>
Full article ">Figure 9
<p>Linearizing control algorithm investigation: (<b>a</b>) control input—glucose feed rate; (<b>b</b>) controlled variable—glucose; (<b>c</b>) biomass concentration; (<b>d</b>) weight of the culture broth.</p>
Full article ">Figure 10
<p>Open-loop control [<a href="#B20-mathematics-12-02205" class="html-bibr">20</a>]. The figure on the left—experimental data for the concentrations of biomass, glucose, and acetate (*), the figure on the right—experimental data for weight, feed rate, oxygen transfer rate (red line), and carbon dioxide transfer rate (blue line).</p>
Full article ">Figure 11
<p>Scheme of the interactive system InSEMCoBio.</p>
Full article ">
14 pages, 2717 KiB  
Review
General ADP-Ribosylation Mechanism Based on the Structure of ADP-Ribosyltransferase–Substrate Complexes
by Hideaki Tsuge, Noriyuki Habuka and Toru Yoshida
Toxins 2024, 16(7), 313; https://doi.org/10.3390/toxins16070313 - 11 Jul 2024
Viewed by 785
Abstract
ADP-ribosylation is a ubiquitous modification of proteins and other targets, such as nucleic acids, that regulates various cellular functions in all kingdoms of life. Furthermore, these ADP-ribosyltransferases (ARTs) modify a variety of substrates and atoms. It has been almost 60 years since ADP-ribosylation [...] Read more.
ADP-ribosylation is a ubiquitous modification of proteins and other targets, such as nucleic acids, that regulates various cellular functions in all kingdoms of life. Furthermore, these ADP-ribosyltransferases (ARTs) modify a variety of substrates and atoms. It has been almost 60 years since ADP-ribosylation was discovered. Various ART structures have been revealed with cofactors (NAD+ or NAD+ analog). However, we still do not know the molecular mechanisms of ART. It needs to be better understood how ART specifies the target amino acids or bases. For this purpose, more information is needed about the tripartite complex structures of ART, the cofactors, and the substrates. The tripartite complex is essential to understand the mechanism of ADP-ribosyltransferase. This review updates the general ADP-ribosylation mechanism based on ART tripartite complex structures. Full article
(This article belongs to the Special Issue ADP-Ribosylation and Beyond)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) ADP-ribosylation proceeds via an oxocarbenium ion after the cleavage of nicotinamide in NAD<sup>+</sup>. (<b>B</b>) ADP-ribosylation targets, specifically amino acids and DNA bases. The purple circles show the atoms that would be modified.</p>
Full article ">Figure 2
<p>(<b>Left</b>) Structure of the C3 exotoxin (green) and the RhoA (GTP) (cyan) complex (PDB: 4XSH). The switch I and switch II regions are shown in red. (<b>Right</b>) A close-up view of the active site, showing 183-GlnXxxGlu-185 in C3 and Asn41 in RhoA. The key glutamate (Glu185) is labeled in red.</p>
Full article ">Figure 3
<p>(<b>Left</b>) Structure of Ia (green) and the actin (cyan) complex (PDB: 4H03). (<b>Right</b>) A close-up view of the active site, showing 378-GluXxxGlu-380 in Ia and Arg177 in actin. The key glutamate (Glu380) is labeled in red.</p>
Full article ">Figure 4
<p>The strain-alleviation model of ADP-ribosylation. Structural changes in the active site and the oxocarbenium ion after cleavage of NAD+ are depicted. Green: immediately after the cleavage of NAD+. Yellow: just before the nucleophilic attack of Arg177. These models were modelled based on the crystal structures of pre- and post-ADP-ribosylation [<a href="#B46-toxins-16-00313" class="html-bibr">46</a>]. The detailed reaction is described in Chapter 2. Glu378 and Glu380 in Ia and Arg 177 and Asp179 in actin are labelled.</p>
Full article ">Figure 5
<p>Structure of the SdeA (green) and ubiquitin (magenta) complex (PDB: 5YIJ). A close-up view of the active site, showing 860-Glu-Xxx-Glu-862 in SdeA and Arg42 (target) and Arg72 in ubiquitin. The key glutamate is labeled in red.</p>
Full article ">Figure 6
<p>Structure of the ScARP (green) and GDP (cyan) complex (PDB: 5ZJ5). (Right) A close-up view of the active site, showing 162-Gln-Xxx-Glu-164 in ScARP and guanine in GDP. The key glutamate (Glu164) is labeled in red.</p>
Full article ">Figure 7
<p>(<b>A</b>) Structures of DarT(T2)–ssDNA complexes (PDB:7OMW and 7ON0). The pre-reaction (green) and post-reaction (ADP-ribosylated) states (cyan) are shown, as are 158-Gln-Xxx-Glu(Glu160Ala)-160 in DarT2, His119, Arg51, and the target thymidine. The key glutamate (Ala160(Glu160Ala)) position is shown using an orange arrow. (<b>B</b>) Structures of DarT1–ssDNA complexes (PDB: 8BAR and 8BAQ). The pre-reaction (green) and post-reaction (ADP-ribosylated) states (purple) are shown, as are 150-Gln-Xxx-Glu(Glu152Ala)-152 in DarT1, Asn104, and the target guanine. The key glutamate (Ala152(Glu152Ala)) position is shown using an orange arrow.</p>
Full article ">
12 pages, 1318 KiB  
Article
The Prediction of Pectin Viscosity Using Machine Learning Based on Physical Characteristics—Case Study: Aglupectin HS-MR
by Przemysław Siejak, Krzysztof Przybył, Łukasz Masewicz, Katarzyna Walkowiak, Ryszard Rezler and Hanna Maria Baranowska
Sustainability 2024, 16(14), 5877; https://doi.org/10.3390/su16145877 - 10 Jul 2024
Viewed by 681
Abstract
In the era of technology development, the optimization of production processes, quality control and at the same time increasing production efficiency without wasting food, artificial intelligence is becoming an alternative tool supporting many decision-making processes. The work used modern machine learning and physical [...] Read more.
In the era of technology development, the optimization of production processes, quality control and at the same time increasing production efficiency without wasting food, artificial intelligence is becoming an alternative tool supporting many decision-making processes. The work used modern machine learning and physical analysis tools to evaluate food products (pectins). Various predictive models have been presented to estimate the viscosity of pectin. Based on the physical analyses, the characteristics of the food product were isolated, including L*a*b* color, concentration, conductance and pH. Prediction was determined using the determination index and loss function for individual machine learning algorithms. As a result of the work, it turned out that the most effective estimation of pectin viscosity was using Decision Tree (R2 = 0.999) and Random Forest (R2 = 0.998). In the future, the prediction of pectin properties in terms of viscosity recognition may be significantly perceived, especially in the food and pharmaceutical industries. Predicting the natural pectin substrate may contribute to improving quality, increasing efficiency and at the same time reducing losses of the obtained final product. Full article
(This article belongs to the Section Sustainable Food)
Show Figures

Figure 1

Figure 1
<p>The variation in dynamic viscosity values as a function of pectin concentration for different pH.</p>
Full article ">Figure 2
<p>The analysis of statistically significantly different viscosity cases: (<b>a</b>)—grouping against pH; (<b>b</b>)—grouping against concentration. a–g: the differences between mean values not sharing same letter were statistically significant (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>A graph of actual values to predicted values for regression models.</p>
Full article ">
17 pages, 9826 KiB  
Article
MeNPs-PEDOT Composite-Based Detection Platforms for Epinephrine and Quercetin
by Sorina Alexandra Leau, Mariana Marin, Ana Maria Toader, Mihai Anastasescu, Cristian Matei, Cecilia Lete and Stelian Lupu
Biosensors 2024, 14(7), 320; https://doi.org/10.3390/bios14070320 - 25 Jun 2024
Cited by 1 | Viewed by 1441
Abstract
The development of low-cost, sensitive, and simple analytical tools for biomolecule detection in health status monitoring is nowadays a growing research topic. Sensing platforms integrating nanocomposite materials as recognition elements in the monitoring of various biomolecules and biomarkers are addressing this challenging objective. [...] Read more.
The development of low-cost, sensitive, and simple analytical tools for biomolecule detection in health status monitoring is nowadays a growing research topic. Sensing platforms integrating nanocomposite materials as recognition elements in the monitoring of various biomolecules and biomarkers are addressing this challenging objective. Herein, we have developed electrochemical sensing platforms by means of a novel fabrication procedure for biomolecule detection. The platforms are based on commercially available low-cost conductive substrates like glassy carbon and/or screen-printed carbon electrodes selectively functionalized with nanocomposite materials composed of Ag and Au metallic nanoparticles and an organic polymer, poly(3,4-ethylenedioxythiophene). The novel fabrication method made use of alternating currents with controlled amplitude and frequency. The frequency of the applied alternating current was 100 mHz for the polymer deposition, while a frequency value of 50 mHz was used for the in situ electrodeposition of Ag and Au nanoparticles. The selected frequency values ensured the successful preparation of the composite materials. The use of readily available composite materials is intended to produce cost-effective analytical tools. The judicious modification of the organic conductive matrix by various metallic nanoparticles, such as Ag and Au, extends the potential applications of the sensing platform toward a range of biomolecules like quercetin and epinephrine, chosen as benchmark analytes for proof-of-concept antioxidant and neurotransmitter detection. The sensing platforms were tested successfully for quercetin and epinephrine determination on synthetic and real samples. Wide linear response ranges and low limit-of-detection values were obtained for epinephrine and quercetin detection. Full article
Show Figures

Figure 1

Figure 1
<p>The sinusoidal current, the potential response, and the changes in frequency recorded for the preparation of PEDOT matrix via the SC approach onto gold-covered quartz crystal electrode by using (<b>a</b>) fixed frequency of 100 mHz and (<b>b</b>) sweeping frequency in the range from 100.1 to 100.001 mHz. Electrolyte solution containing 10<sup>−2</sup> M EDOT in 0.1 M LiClO<sub>4</sub>.</p>
Full article ">Figure 2
<p>Two-dimensional enhanced-contrast topographic AFM images (with characteristic line scans) scanned over 10 µm × 10 µm of (<b>a</b>) PEDOT, (<b>b</b>) PEDOT-AgNPs, (<b>c</b>) PEDOT-AuNPs; SEM images of (<b>d</b>) PEDOT, (<b>e</b>) PEDOT-AgNPs, and (<b>f</b>) PEDOT-AuNPs. EDX elemental mapping and spectra for (<b>g</b>,<b>i</b>) PEDOT-AgNPs, and (<b>h</b>,<b>j</b>) PEDOT-AuNPs.</p>
Full article ">Figure 3
<p>(<b>a</b>) CV traces registered at unmodified GCE, PEDOT, and PEDOT-AgNPs at 50 mV/s potential scanning rate in 5 mM Na<sub>4</sub>[Fe(CN)<sub>6</sub>], with 0.5 M KNO<sub>3</sub> as supporting electrolyte. (<b>b</b>) The corresponding EIS spectra of bare GCE, PEDOT, and PEDOT-AgNPs recorded in 5 mM Na<sub>4</sub>[Fe(CN)<sub>6</sub>]/K<sub>3</sub>[Fe(CN)<sub>6</sub>], with 0.5 M KNO<sub>3</sub> as supporting electrolyte. Inset: the equivalent electrical circuit used in the fitting of the experimental data.</p>
Full article ">Figure 4
<p>(<b>a</b>) CV registered with GCE, PEDOT, and PEDOT-AgNPs sensing platform in acetate buffer system (pH = 5) with 40 μM QR, with a 50 mV/s potential sweep rate. (<b>b</b>) Influence of the pH on the QR detection of the PEDOT-AgNPs sensing platform in buffered systems with pH ranging from 3 to 8 containing 50 μM QR. Potential sweep rate: 50 mV/s.</p>
Full article ">Figure 5
<p>(<b>a</b>) CVs for the PEDOT-AgNPs sensing platform in acetate buffer system (pH = 5) in the presence of various amounts of QR: 1, 5, 10, 15, 20, 30, and 40 μM. Potential sweep rate: 50 mV/s. (<b>b</b>) The calibration curve of the AgNPs platform.</p>
Full article ">Figure 6
<p>CVs for the PEDOT-AgNPs platform in acetate buffer system (pH = 5) containing QR and various amounts of (<b>a</b>) cysteine and (<b>b</b>) ascorbic acid. (<b>c</b>) The response of the PEDOT-AgNPs platform in diluted grape must sample spiked with 20, 25, and 30 μM QR in acetate buffer system (pH = 5), with a 50 mV/s potential scanning rate.</p>
Full article ">Figure 7
<p>(<b>a</b>) Analytical response of the PEDOT-AuNPs sensing platform in phosphate buffer (pH = 7) containing various amounts of EPI (1, 2, 4, 6, 8, 10, 20, 40, 60, 80, and 100 μM); 50 mV/s potential scan rate. Inset: the corresponding calibration plot. (<b>b</b>) Analytical response of the PEDOT-AuNPs platform in phosphate buffer (pH = 7) containing 40 μM EPI and various amounts of DA: 10, 20, 60, 100, 200, and 300 μM. (<b>c</b>) Analytical response of the AuNPs-based platform in urine sample spiked with various amounts of EPI: 6, 10, 20, and 40 μM. (<b>d</b>) Analytical response of the AuNPs-based sensing platform in urine sample spiked with various amounts of UA of 60, 120, 180 μM, and 10 μM EPI, with a 50 mV/s potential scan rate.</p>
Full article ">Scheme 1
<p>The structure of quercetin and its oxidized tris-ketone forms.</p>
Full article ">
28 pages, 5831 KiB  
Article
The RNA Helicase Ded1 from Yeast Is Associated with the Signal Recognition Particle and Is Regulated by SRP21
by Hilal Yeter-Alat, Naïma Belgareh-Touzé, Agnès Le Saux, Emmeline Huvelle, Molka Mokdadi, Josette Banroques and N. Kyle Tanner
Molecules 2024, 29(12), 2944; https://doi.org/10.3390/molecules29122944 - 20 Jun 2024
Cited by 1 | Viewed by 761
Abstract
The DEAD-box RNA helicase Ded1 is an essential yeast protein involved in translation initiation that belongs to the DDX3 subfamily. The purified Ded1 protein is an ATP-dependent RNA-binding protein and an RNA-dependent ATPase, but it was previously found to lack substrate specificity and [...] Read more.
The DEAD-box RNA helicase Ded1 is an essential yeast protein involved in translation initiation that belongs to the DDX3 subfamily. The purified Ded1 protein is an ATP-dependent RNA-binding protein and an RNA-dependent ATPase, but it was previously found to lack substrate specificity and enzymatic regulation. Here we demonstrate through yeast genetics, yeast extract pull-down experiments, in situ localization, and in vitro biochemical approaches that Ded1 is associated with, and regulated by, the signal recognition particle (SRP), which is a universally conserved ribonucleoprotein complex required for the co-translational translocation of polypeptides into the endoplasmic reticulum lumen and membrane. Ded1 is physically associated with SRP components in vivo and in vitro. Ded1 is genetically linked with SRP proteins. Finally, the enzymatic activity of Ded1 is inhibited by SRP21 in the presence of SCR1 RNA. We propose a model where Ded1 actively participates in the translocation of proteins during translation. Our results provide a new understanding of the role of Ded1 during translation. Full article
Show Figures

Figure 1

Figure 1
<p>Secondary structure model of yeast SCR1 from Zwieb et al. [<a href="#B32-molecules-29-02944" class="html-bibr">32</a>]. The model is based on phylogenetically conserved features found in SRP RNAs and on structural probing experiments [<a href="#B32-molecules-29-02944" class="html-bibr">32</a>,<a href="#B33-molecules-29-02944" class="html-bibr">33</a>]. Yeast and other fungal SRP RNAs are unusual in that they are much larger than in other organisms, and they lack the characteristic structure consisting of hairpins 3 and 4 of the Alu domain. Yeast has additional hairpins 9, 10, 11, and 12 that are poorly characterized and that have other proposed secondary structures. Conserved sequence motifs and tertiary interactions are shown in gray.</p>
Full article ">Figure 2
<p>Ded1-IgG pull-downs of yeast extracts. Ded1-specific IgG (Ded1-IgG) or IgG from pre-immune serum (Pre-IgG) were used to recover the associated factors. Input, a fraction of the yeast extract used in the pull-down experiments was directly loaded onto the gel or RT-PCR amplified. (<b>A</b>) Purified RNA from yeast extracts (~20% of input) or from IgG pull-downs was reverse transcribed and PCR amplified for 25 cycles with gene-specific oligonucleotides. The resulting products were electrophoretically separated on a 2% agarose gel containing ethidium bromide, and the products were visualized with a Gel Doc XR+ (Bio-Rad, Hercules, CA, USA). (<b>B</b>) Western blot analysis of HA-tagged SRP proteins. Proteins were electrophoretically separated on a 12% SDS-PAGE, transferred to nitrocellulose membranes, and then revealed with anti-HA IgG. Input, 40 µg (~10%) of the yeast extract was directly loaded on the gel. Ded1-IgG, Ded1-specific IgG was used to pull down Ded1-associated proteins. IgG+RNase, complexes bound to Ded1-IgG-protein A beads were digested with RNase A (1 mg/mL) prior to washing and elution. Note that the same exposure was used for the input, Ded1-IgG, and pre-IgG in panel A, while somewhat shorter exposures were used for the input in panel B.</p>
Full article ">Figure 3
<p>Ded1 multicopy suppression of SRP protein depletions. Cells of the indicated strains with the <span class="html-italic">TET</span> promoter were grown in SD-LEU medium, serially diluted by a factor of 10, and spotted on SD-LEU agar plates with (+DOX) or without (−DOX) 10 µg/mL of doxycycline. Cultures were grown for 4 days at 36 °C. p415, empty <span class="html-italic">LEU</span> plasmid; <span class="html-italic">GPD-DED1</span>, Ded1 in p415 with the high-expression <span class="html-italic">GPD</span> promoter and <span class="html-italic">CYC1</span> terminator; <span class="html-italic">GPD-F162C</span>, a Ded1 mutant with reduced ATP binding and enzymatic activity [<a href="#B48-molecules-29-02944" class="html-bibr">48</a>]. BY4742, a wildtype yeast strain showing unimpeded growth. The phenotypes were most apparent at 36 °C, but similar effects were obtained at 30 °C.</p>
Full article ">Figure 4
<p>Cellular location of Ded1 relative to the ER and SRP proteins. (<b>A</b>) Ded1-DQAD-GFP was expressed in the <span class="html-italic">sec62</span> temperature-sensitive mutant with the integrated <span class="html-italic">KAR2-RFP</span> plasmid and grown to an OD600 of 1.0 at 24 °C. (<b>B</b>) The same cells as in A were incubated for 15 min at the non-permissive temperature of 37 °C prior to visualization. The arrowheads indicate positions where chains of Ded1-DQAD-GFP foci co-associated with Kar2-RFP in panels B and C. (<b>C</b>) SRP14-GFP expressed from the chromosome and Ded1-DQAD-mCh expressed off the p415 plasmid were grown to an OD600 of 0.95 at 30 °C. (<b>D</b>) SRP14-GFP was overexpressed off the p413-PL plasmid and Ded1-DQAD-mCh was overexpressed off the p416-PL plasmid until an OD600 of 0.4 at 30 °C in the xpoI-T539C yeast strain. Cells in the insert were treated with 10 µg/µL (~200 nM) of leptomycin b for 1 h. The arrowheads indicate the positions where Ded1-DQAD and SRP14 colocalize within a crescent-shaped region of the nucleus.</p>
Full article ">Figure 5
<p>Overexpressed SRP14 and SRP21 accumulate in the nucleus. (<b>A</b>) SRP14-GFP was expressed off the p413 plasmid and Ded1-DQAD-mCh was expressed off the p416 plasmid in the <span class="html-italic">xpoI-T539C</span> yeast strain [<a href="#B72-molecules-29-02944" class="html-bibr">72</a>] and grown to an OD600 of 0.45 at 30 °C. (<b>B</b>) The same as in A except that the cells were incubated for 60 min in the presence of 10 µg/mL of leptomycin b. (<b>C</b>) Same as A but with cells expressing SRP21-GFP. The arrowheads indicate positions where SRP21 form nuclear foci. (<b>D</b>) Same as C but with cells incubated for 60 min with leptomycin b.</p>
Full article ">Figure 6
<p>Ded1 physically interacted with the SRP proteins in the absence of RNA. (<b>A</b>) A total of 4 µg of Ded1 was incubated with 4 µg of each SRP protein. The material was incubated for 45 min at 30 °C, immunoprecipitated with protein-A-Sepharose beads with Ded1-specific IgG, separated on a 12% SDS PAGE, and visualized with Coomassie blue. SRP68 and SRP72 migrated close to Ded1 and were consequently not unambiguously separated. (<b>B</b>) The same as A except 6 µg of the SRP proteins was used with 4 µg of Ded1. Proteins were recovered with Ded1- or SRP21-specific IgG as indicated.</p>
Full article ">Figure 7
<p>The SRP proteins inhibited the ATPase activity of Ded1. (<b>A</b>) Reactions were undertaken with 7 nM of Ded1, 200 nM of the SRP proteins, 1 mM of ATP, and 23 nM of SCR1 or actin RNA. The reaction velocities were measured over 40 min at 30 °C. (<b>B</b>) Reactions were conducted as in A but with 23 nM of the SRP proteins except for SRP14, which was used at 46 nM to form the homodimer, and 23 nM of RNAs. The reaction velocities were normalized relative to the activity of Ded1 in the presence of the RNA (SCR1 or actin) alone. +SRP-All, Ded1 was incubated with SRP14, SRP21, SRP54, Sec65, SRP68, and SRP72; Ded1-GAT, a Ded1 phosphate-binding (P-loop) mutant that lacks ATPase activity; +no RNA, Ded1 was incubated in the absence of an RNA substrate with the SRP proteins. The means and standard deviations are shown for two independent experiments in panel (<b>A</b>) and for three in panel (<b>B</b>). The lower error bars were deleted for clarity.</p>
Full article ">Figure 8
<p>The RNA-dependent effects of SRP21 on the ATPase activity of Ded1. (<b>A</b>) Ded1 was pre-incubated with the RNAs at 30 °C for 30 min. SRP21 or SRP21∆73 was then added at 200 nM with 1 mM of ATP, and the ATPase velocity was measured over 40 min. The means and standard deviations are shown for two independent experiments. (<b>B</b>) Ded1 at 7 nM was incubated with 23 nM of SCR1, 23 nM of actin, or 0.14 µg/µL of yeast RNA and with 1 mM of ATP. SRP21 was used at 23 nM and SRP14 at 46 nM (to form a homodimer). The reaction velocities were measured over 40 min at 30 °C. The means and standard deviations are shown for three independent measurements for SCR1 and actin and for two independent measurements for yeast RNA. (<b>C</b>) Reactions were performed as in B. Ded1 at 7 nM was incubated with 23 nM of SCR1 (equivalent to 0.0039 µg/µL) or with 0.12 µg/µL of tRNA or poly(A). SRP21 was used at 200 nM. The means and standard deviations are shown for three independent measurements. The lower error bars were deleted for clarity.</p>
Full article ">Figure 9
<p>RNA-binding assays of Ded1 and SRP21. (<b>A</b>) Ded1 binds SCR1 and actin with similar affinities. The indicated quantities of the Ded1 protein were incubated with 0.15 µM of the indicated RNAs and then separated on a 1% agarose gel in the presence of ethidium bromide. Ori, loading well of agarose gel. (<b>B</b>) The indicated quantities of the SRP21 proteins were incubated with 50 nM of the indicated RNAs and separated on a 1% agarose gel. (<b>C</b>) SRP21 was deleted for the 73 carboxyl-terminal residues that are not structurally conserved in mammalian SRP9.</p>
Full article ">Figure 10
<p>Model for the role of Ded1 in SRP-dependent translation. (<b>A</b>) Ded1 (shown in green) associates with the mRNA during translation initiation and remains attached to the mRNA in front of the ribosomes. It consists of RecA-like domains 1 (d1) and 2 (d2), an amino-terminal domain (N), and a carboxyl-terminal domain (C). The RNA-dependent ATPase activity of Ded1 is unaltered, and it is often in the “open” conformation with weak affinity for the RNA; it is able to translocate with the ribosomes during translation. The interactions with the 3′-bound Pab2 are not indicated in this cartoon. (<b>B</b>) The SRP (shown in blue) associates with ribosomes translating mRNAs (or undergoes conformational changes in the case of pre-bound SRP) when the signal peptide leaves the exit channel and obtains a certain length. Ded1 may help in assembling and stabilizing the complex. Conformational changes of the SRP cause SRP14 to block the entry channel and prevent the eEF2 elongation factor (EF) from binding the ribosomes, which pauses elongation. Ded1 may bind part of the Alu domain of SCR1, shown in magenta, during these conformational changes to promote SRP14 binding to the ribosomes. At the same time, SRP21 inhibits the ATPase activity of Ded1, which forms the “closed” conformation with a high affinity for the RNA. This ATP-bound form of Ded1 kinks the RNA (red triangle) on domain 1 and locks Ded1 on the RNA. This prevents the ribosomes from frameshifting (sliding) on the RNA and perhaps stabilizes the ribosome-mRNA complex to prevent the premature termination of translation. (<b>C</b>) The paused mRNA-ribosome complex associates with SRP receptor (SR) factors SRP101 and subsequently SRP102, which brings the mRNA-ribosome complex to the Sec61 ER translocon. (<b>D</b>) The SRP complex dissociates from the ribosomes, the ATPase activity of Ded1 is restored, and translation continues. Note that this model also applies to the SRP-dependent import of polypeptides with internal transmembrane domains, and it does not preclude the possibility that multiple Ded1 molecules are involved, that the SRP associates multiple times with the ribosomes during elongation, or that the SRP-associated ribosomes remain on the ER over multiple rounds of translation.</p>
Full article ">
17 pages, 4341 KiB  
Review
Unveiling the Druggable Landscape of Bacterial Peptidyl tRNA Hydrolase: Insights into Structure, Function, and Therapeutic Potential
by Surbhi Mundra and Ashish Kabra
Biomolecules 2024, 14(6), 668; https://doi.org/10.3390/biom14060668 - 7 Jun 2024
Viewed by 1111
Abstract
Bacterial peptidyl tRNA hydrolase (Pth) or Pth1 emerges as a pivotal enzyme involved in the maintenance of cellular homeostasis by catalyzing the release of peptidyl moieties from peptidyl-tRNA molecules and the maintenance of a free pool of specific tRNAs. This enzyme is vital [...] Read more.
Bacterial peptidyl tRNA hydrolase (Pth) or Pth1 emerges as a pivotal enzyme involved in the maintenance of cellular homeostasis by catalyzing the release of peptidyl moieties from peptidyl-tRNA molecules and the maintenance of a free pool of specific tRNAs. This enzyme is vital for bacterial cells and an emerging drug target for various bacterial infections. Understanding the enzymatic mechanisms and structural intricacies of bacterial Pth is pivotal in designing novel therapeutics to combat antibiotic resistance. This review provides a comprehensive analysis of the multifaceted roles of Pth in bacterial physiology, shedding light on its significance as a potential drug target. This article delves into the diverse functions of Pth, encompassing its involvement in ribosome rescue, the maintenance of a free tRNA pool in bacterial systems, the regulation of translation fidelity, and stress response pathways within bacterial systems. Moreover, it also explores the druggability of bacterial Pth, emphasizing its promise as a target for antibacterial agents and highlighting the challenges associated with developing specific inhibitors against this enzyme. Structural elucidation represents a cornerstone in unraveling the catalytic mechanisms and substrate recognition of Pth. This review encapsulates the current structural insights of Pth garnered through various biophysical techniques, such as X-ray crystallography and NMR spectroscopy, providing a detailed understanding of the enzyme’s architecture and conformational dynamics. Additionally, biophysical aspects, including its interaction with ligands, inhibitors, and substrates, are discussed, elucidating the molecular basis of bacterial Pth’s function and its potential use in drug design strategies. Through this review article, we aim to put together all the available information on bacterial Pth and emphasize its potential in advancing innovative therapeutic interventions and combating bacterial infections. Full article
(This article belongs to the Special Issue The Structure and Function of Proteins, Lipids and Nucleic Acids)
Show Figures

Figure 1

Figure 1
<p>Multiple sequence alignment of Pth proteins from <span class="html-italic">E. coli</span>, <span class="html-italic">P. aeruginosa</span>, <span class="html-italic">S. aureus</span>, <span class="html-italic">S. cerevisiae</span>, <span class="html-italic">S. solfataricus</span>, <span class="html-italic">M. tuberculosis</span>, and <span class="html-italic">H. sapiens</span> is shown.</p>
Full article ">Figure 2
<p>The release and cycling of peptidyl tRNA: The incomplete translation leads to the release of peptidyl tRNAs in the cell. The Pth protein enzymatically cleaves these peptidyl tRNAs, thus releasing free tRNAs in the cell for further rounds of protein synthesis. This figure was created with <a href="https://www.biorender.com/" target="_blank">https://www.biorender.com/</a> (accessed on 1 June 2024).</p>
Full article ">Figure 3
<p>A timeline of the advancement in the knowledge of peptidyl tRNA hydrolase (Pth) [<a href="#B7-biomolecules-14-00668" class="html-bibr">7</a>,<a href="#B10-biomolecules-14-00668" class="html-bibr">10</a>,<a href="#B15-biomolecules-14-00668" class="html-bibr">15</a>,<a href="#B16-biomolecules-14-00668" class="html-bibr">16</a>,<a href="#B17-biomolecules-14-00668" class="html-bibr">17</a>,<a href="#B18-biomolecules-14-00668" class="html-bibr">18</a>,<a href="#B19-biomolecules-14-00668" class="html-bibr">19</a>,<a href="#B20-biomolecules-14-00668" class="html-bibr">20</a>].</p>
Full article ">Figure 4
<p><span class="html-italic">E. coli</span> Pth structure (PDB ID: 2PTH) showing structural organization in bacterial Pth: <b>Top panel</b> is the cartoon representation of the protein which shows the arrangement of the helices and sheets in the tertiary structure of Pth (forming a closed cylindrical structure), with the base loop, lid loop, and the gate loop colored red, yellow, and blue, respectively. The <b>lower panel</b> shows the surface representation of the protein with all three loop regions, surrounding the active site crevice on the surface. The structures on the right side of the figure are the flipped view orientation of the protein by 90°.</p>
Full article ">Figure 5
<p>Cartoon representation of Pth proteins with labeled helices and sheets: (<b>Left</b>) Pth1 from <span class="html-italic">E. coli</span> (2PTH) comprising six alpha helices and seven beta strands, and (<b>Right</b>) Pth2 from <span class="html-italic">H. sapiens</span> (1Q7S) with four alpha helices and four beta strands.</p>
Full article ">Figure 6
<p>Representation of Pth’s action on its substrate N-blocked aminoacyl-tRNA: The ester bond between tRNA and peptide or N-blocked aminoacyl-tRNA (Green color represents the blockage of N-terminal, which can either be a fluorophore/molecule blocking the N-terminus of the peptide or a linked amino acid chain) is cleaved after the Pth action and free tRNA and N-blocked amino acid/peptide are released.</p>
Full article ">Figure 7
<p>The orientation of active site residues in bacterial Pth: The active site residues N10, H20, D93, N68, and N114 form a crevice at the surface of the protein and shown in cyan color. Two aromatic residues Y15 and F66 are shown in green color that helps in stacking of the substrate whereas K142 of lid loop is shown in yellow color that interacts with CCA terminus of tRNA moiety. A zoomed-in view of the organization of the residues is shown on the right.</p>
Full article ">Figure 8
<p>Overlay of Pth proteins: (<b>Left</b>) Pth1 proteins from <span class="html-italic">E. coli</span> (green), <span class="html-italic">A. baumannii</span> (cyan), <span class="html-italic">S. aureus</span> (orange), <span class="html-italic">K. pneumoniae</span> (yellow), <span class="html-italic">P. aeruginosa</span> (red) and <span class="html-italic">V. cholerae</span> (pink). The overlay shows the structural similarity among Pth1 from these different organisms. (<b>Right</b>) Pth2 from <span class="html-italic">S. sulfotaricus</span> (blue), <span class="html-italic">H. sapiens</span> (orange), <span class="html-italic">A. fulgidis</span> (green) and <span class="html-italic">T. acidophilum</span> (beige).</p>
Full article ">
19 pages, 3128 KiB  
Article
Probing the Conformational Restraints of DNA Damage Recognition with β-L-Nucleotides
by Anna V. Yudkina, Daria V. Kim, Timofey D. Zharkov, Dmitry O. Zharkov and Anton V. Endutkin
Int. J. Mol. Sci. 2024, 25(11), 6006; https://doi.org/10.3390/ijms25116006 - 30 May 2024
Viewed by 673
Abstract
The DNA building blocks 2′-deoxynucleotides are enantiomeric, with their natural β-D-configuration dictated by the sugar moiety. Their synthetic β-L-enantiomers (βLdNs) can be used to obtain L-DNA, which, when fully substituted, is resistant to nucleases and is finding use in many biosensing and nanotechnology [...] Read more.
The DNA building blocks 2′-deoxynucleotides are enantiomeric, with their natural β-D-configuration dictated by the sugar moiety. Their synthetic β-L-enantiomers (βLdNs) can be used to obtain L-DNA, which, when fully substituted, is resistant to nucleases and is finding use in many biosensing and nanotechnology applications. However, much less is known about the enzymatic recognition and processing of individual βLdNs embedded in D-DNA. Here, we address the template properties of βLdNs for several DNA polymerases and the ability of base excision repair enzymes to remove these modifications from DNA. The Klenow fragment was fully blocked by βLdNs, whereas DNA polymerase κ bypassed them in an error-free manner. Phage RB69 DNA polymerase and DNA polymerase β treated βLdNs as non-instructive but the latter enzyme shifted towards error-free incorporation on a gapped DNA substrate. DNA glycosylases and AP endonucleases did not process βLdNs. DNA glycosylases sensitive to the base opposite their cognate lesions also did not recognize βLdNs as a correct pairing partner. Nevertheless, when placed in a reporter plasmid, pyrimidine βLdNs were resistant to repair in human cells, whereas purine βLdNs appear to be partly repaired. Overall, βLdNs are unique modifications that are mostly non-instructive but have dual non-instructive/instructive properties in special cases. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Structures of normal β-D- (top) and mirror-image β-L-deoxynucleotides (bottom). (<b>b</b>) Structure of normal DNA (top, 5′-TTC-3′/5′-GAA-3′, PDB ID 355D [<a href="#B12-ijms-25-06006" class="html-bibr">12</a>]) and a hypothetical structure of 5′-T[βLdT]C-3′/5′-GAA-3′ maintaining Watson–Crick bonding in DNA (bottom; see <a href="#sec4-ijms-25-06006" class="html-sec">Section 4</a> for the description of model building). Dashes indicate hydrogen bonds.</p>
Full article ">Figure 2
<p>Primer extension by RBpol and Pol κ across βLdNs on a primer–template substrate. (<b>a</b>) Scheme of the substrate. FAM, fluorescent label; β, modified nucleotide. (<b>b</b>) Nucleotide incorporation by RBpol. (<b>c</b>) Nucleotide incorporation by Pol κ. +1 …+3, number of nucleotides added.</p>
Full article ">Figure 3
<p>Primer extension by Pol β across βLdNs on primer–template(pt) and gapped (gap) substrates. (<b>a</b>) Scheme of the substrates. FAM, fluorescent label; β, modified nucleotide; p, 5′-terminal phosphate. (<b>b</b>–<b>e</b>) Insertion of dNMPs by Pol β opposite to βLdA (<b>b</b>), βLdG (<b>c</b>), βLdT (<b>d</b>) and βLdC (<b>e</b>). +1 …+2, number of nucleotides added.</p>
Full article ">Figure 4
<p>Cleavage of substrates containing βLdNs by DNA glycosylases MutY (<b>a</b>), MBD4 (<b>b</b>), Fpg (<b>c</b>) and OGG1 (<b>d</b>). S, oligonucleotide substrate; P, cleavage product. βA, βLdA; βC, βLdC; βG, βLdG; βT, βLdT.</p>
Full article ">Figure 5
<p>Cleavage of substrates containing βLdNs by AP endonucleases. (<b>a</b>) Cleavage of βLdA:T, βLdT:A, βLdG:C and βLdC:G substrates by human APE1 and <span class="html-italic">E. coli</span> Nfo and Xth. (<b>b</b>) Cleavage of βLdA:T by Xth under different conditions. S, substrate, P, cleavage product.</p>
Full article ">Figure 6
<p>Transcriptional mutagenesis and repair induced by βLdNs in HeLa cells. (<b>a</b>) Scheme of the processes induced by a lesion in the <span class="html-italic">eGFP</span> reporter gene. (<b>b</b>) Relative EGFP expression normalized for the fluorescence of the control G-construct (<span class="html-italic">n</span> = 3, mean ± SD shown). Differences between constructs: <span class="html-italic">p</span> &lt; 0.05 (*); <span class="html-italic">p</span> &lt; 0.01 (**); <span class="html-italic">p</span> &lt; 0.005 (***); two-tailed Student’s <span class="html-italic">t</span>-test with Bonferroni correction applied. The brightness of the bars is proportional to the EGFP expression level.</p>
Full article ">
Back to TopTop