Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where Qingfeng Pan is active.

Publication


Featured researches published by Qingfeng Pan.


The Journal of Experimental Biology | 2005

Comparative overwintering physiology of Alaska and Indiana populations of the beetle Cucujus clavipes (Fabricius): roles of antifreeze proteins, polyols, dehydration and diapause.

Valerie A. Bennett; Todd Sformo; Kent R. Walters; Øivind Tøien; Kennan Jeannet; Ronald Hochstrasser; Qingfeng Pan; Anthony S. Serianni; Brian M. Barnes; John G. Duman

SUMMARY The beetle Cucujus clavipes is found in North America over a broad latitudinal range from North Carolina (latitude ∼35°N) to near tree line in the Brooks Range in Alaska (latitude, ∼67°30′ N). The cold adaptations of populations from northern Indiana (∼41°45′ N) and Alaska were compared and, as expected, the supercooling points (the temperatures at which they froze) of these freeze-avoiding insects were significantly lower in Alaska insects. Both populations produce glycerol, but the concentrations in Alaska larvae were much higher than in Indiana insects (∼2.2 and 0.5 mol l–1, respectively). In addition, both populations produce antifreeze proteins. Interestingly, in the autumn both populations have the same approximate level of hemolymph thermal hysteresis, indicative of antifreeze protein activity, suggesting that they synthesize similar amounts of antifreeze protein. A major difference is that the Alaska larvae undergo extreme dehydration in winter wherein water content decreases from 63–65% body water (1.70–1.85 g H2O g–1 dry mass) in summer to 28–40% body water (0.40–0.68 g H2O g–1 dry mass) in winter. These 2.5–4.6-fold reductions in body water greatly increase the concentrations of antifreeze in the Alaska insects. Glycerol concentrations would increase to 7–10 mol l–1 while thermal hysteresis increased to nearly 13°C (the highest ever measured in any organism) in concentrated hemolymph. By contrast, Indiana larvae do not desiccate in winter. The Alaska population also undergoes a diapause while insects from Indiana do not. The result of these, and likely additional, adaptations is that while the mean winter supercooling points of Indiana larvae were approximately –23°C, those of Alaska larvae were –35 to– 42°C, and at certain times Alaska C. clavipes did not freeze when cooled to –80°C.


Journal of Biological Chemistry | 2009

Cryoprotectant Biosynthesis and the Selective Accumulation of Threitol in the Freeze-tolerant Alaskan Beetle, Upis ceramboides

Kent R. Walters; Qingfeng Pan; Anthony S. Serianni; John G. Duman

Adult Upis ceramboides do not survive freezing in the summer but tolerate freezing to −60 °C in midwinter. The accumulation of two cryoprotective polyols, sorbitol and threitol, is integral to the extraordinary cold-hardiness of this beetle. U. ceramboides are the only animals known to accumulate high concentrations of threitol; however, the biosynthetic pathway has not been studied. A series of 13C-labeled compounds was employed to investigate this biosynthetic pathway using 13C{1H} NMR spectroscopy. In vivo metabolism of 13C-labeled glucose isotopomers demonstrates that C-3—C-6 of glucose become C-1—C-4 of threitol. This labeling pattern is expected for 4-carbon saccharides arising from the pentose phosphate pathway. In vitro experiments show that threitol is synthesized from erythrose 4-phosphate, a C4 intermediate in the PPP. Erythrose 4-phosphate is epimerized and/or isomerized to threose 4-phosphate, which is subsequently reduced by a NADPH-dependent polyol dehydrogenase and dephosphorylated by a sugar phosphatase to form threitol. Threitol 4-phosphate appears to be the preferred substrate of the sugar phosphatase(s), promoting threitol synthesis over that of erythritol. In contrast, the NADPH-dependent polyol dehydrogenase exhibits broad substrate specificity. Efficient erythritol catabolism under conditions that promote threitol synthesis, coupled with preferential threitol biosynthesis, appear to be responsible for the accumulation of high concentrations of threitol (250 mm) without concomitant accumulation of erythritol.


Acta Crystallographica Section C-crystal Structure Communications | 2006

Methyl 4-O-β-l-galactopyranosyl-β-d-glucopyranoside (methyl β-l-lactoside)

Qingfeng Pan; Bruce C. Noll; Anthony S. Serianni

Methyl beta-L-lactoside, C13H24O11, (II), is described by glycosidic torsion angles phi (O5Gal-C1Gal-O4Glc-C4Glc) and psi (C1Gal-O1Gal-C4Glc-C5Glc) of 93.89 (13) and -127.43 (13) degrees , respectively, where the ring atom numbering conforms to the convention in which C1 is the anomeric C atom and C6 is the exocyclic hydroxymethyl (CH2OH) C atom in both residues (Gal is galactose and Glc is glucose). Substitution of L-Gal for D-Gal in the biologically relevant disaccharide, methyl beta-lactoside [Stenutz, Shang & Serianni (1999). Acta Cryst. C55, 1719-1721], (I), significantly alters the glycosidic linkage interface. In the crystal structure of (I), one inter-residue (intramolecular) hydrogen bond is observed between atoms H3OGlc and O5Gal. In contrast, in the crystal structure of (II), inter-residue hydrogen bonds are observed between atoms H6OGlc and O5Gal, H6OGlc and O6Gal, and H3OGlc and O2Gal, with H6OGlc serving as a donor with two intramolecular acceptors.


Acta Crystallographica Section C-crystal Structure Communications | 2005

Methyl 4-O-β-D-galactopyranosyl α-D-glucopyranoside (methyl α-lactoside)

Qingfeng Pan; Bruce C. Noll; Anthony S. Serianni

Methyl alpha-lactoside, C13H24O11, (I), is described by glycosidic torsion angles varphi (O5gal-C1gal-O1gal-C4glc) and psi (C1gal-O1gal-C4glc-C5glc), which have values of -93.52 (13) and -144.83 (11) degrees, respectively, where the ring atom numbering conforms to the convention in which C1 is the anomeric C atom and C6 is the exocyclic hydroxymethyl (-CH2OH) C atom in both residues. The linkage geometry is similar to that observed in methyl beta-lactoside methanol solvate, (II), in which varphi is -88.4 (4) degrees and psi is -161.3 (4) degrees. As in (II), an intermolecular O3glc-H...O5gal hydrogen bond is observed in (I). The hydroxymethyl group conformation in both residues is gauche-trans, with torsion angles omegagal (O5gal-C5gal-C6gal-O6gal) and omega(glc) (O5glc-C5glc-C6glc-O6glc) of 69.15 (13) and 72.55 (14) degrees, respectively. The latter torsion angle differs substantially from that found for (II) [-54.6 (2) degrees; gauche-gauche]. Cocrystallization of methanol, which is hydrogen bonded to O6glc in the crystal structure of (II), presumably affects the hydroxymethyl conformation in the Glc residue in (II).


Tetrahedron | 2017

Rapid assembly of branched mannose oligosaccharides through consecutive regioselective glycosylation: A convergent and efficient strategy

Bo Meng; Jun Wang; Qianli Wang; Anthony S. Serianni; Qingfeng Pan

A convergent and efficient strategy for the synthesis of high-mannose oligosaccharides is described wherein regioselective glycosylations between trichloroacetimidate donors and partially protected acceptors are employed to reduce the number of protection-deprotection steps. Two representative branched mannose oligosaccharides, a mannose heptasaccharide (Man7) and a mannose nonasaccharide (Man9) were constructed via (4+3) and (5+4) glycosylations, respectively. These mannose-containing oligosaccharides were obtained in nine steps in ~25% overall yield and >98% purity on 60-70 mg scales to demonstrate the effectiveness of the strategy.


Journal of Labelled Compounds and Radiopharmaceuticals | 2016

A chemical synthesis of a multiply 13C‐labeled hexasaccharide: a high‐mannose N‐glycan fragment

Wenhui Zhang; Qingfeng Pan; Anthony S. Serianni

As covalent modifiers of proteins, high-mannose N-glycans are important in maintaining protein structure and function in vivo. The conformations of these glycans can be studied by nuclear magnetic resonance spectroscopy using spin-spin couplings (J-couplings; scalar couplings) and other nuclear magnetic resonance parameters that are sensitive to the geometries of their constituent glycosidic linkages and other mobile elements in their structures. These analyses often require 13 C-labeling at specific carbon atoms, especially when measurements of 13 C-13 C J-couplings are of interest. The selection of particular 13 C isotopomers of a glycan depends on the type of question under scrutiny. A chemical synthesis of a mannose-containing hexasaccharide, α[1-13 C]Man(1→2)α[1,2-13 C2 ]Man(1→6)[α[1-13 C]Man(1→2)α[1,2-13 C2 ]Man(1→3)]α[1,2-13 C2 ]Man(1→6)βManOCH3 , which is a nested fragment of the high-mannose N-glycans of human glycoproteins and contains eight 13 C-enriched carbon sites, is described in this report. The selected 13 C isotopomer was chosen to maximize the measurement of J-couplings sensitive to linkage conformations. This work demonstrates that chemical syntheses of multiply 13 C-labeled oligosaccharides are technically feasible and practical using present synthetic methods. The availability of this and other multiply 13 C-labeled mannose-containing oligosaccharides will promote future studies of their conformations in solution and in the bound state.


Carbohydrate Research | 2018

Synthesis of high-mannose oligosaccharides containing mannose-6-phosphate residues using regioselective glycosylation

Bo Meng; Jun Wang; Quanli Wang; Anthony S. Serianni; Qingfeng Pan

Molecular recognition of mannose-6-phosphate (M6P)-modified oligosaccharides by transmembrane M6P receptors is a key signaling event in lysosomal protein trafficking in vivo. Access to M6P-containing high-mannose N-glycans is essential to achieving a thorough understanding of the M6P ligand-receptor recognition process. Herein we report the application of a versatile and reliable chemical strategy to prepare asymmetric di-antennary M6P-tagged high-mannose oligosaccharides in >20% overall yield and in high purity (>98%). Regioselective chemical glycosylation coupled with effective phosphorylation and product purification protocols were applied to rapidly assemble these oligosaccharides. The development of this synthetic strategy simplifies the preparation of M6P-tagged high-mannose oligosaccharides, which will improve access to these compounds to study their structures and biological functions.


Acta Crystallographica Section C-crystal Structure Communications | 2005

Methyl 4-O-beta-D-galactopyranosyl alpha-D-glucopyranoside (methyl alpha-lactoside).

Qingfeng Pan; Bruce C. Noll; Anthony S. Serianni

Methyl alpha-lactoside, C13H24O11, (I), is described by glycosidic torsion angles varphi (O5gal-C1gal-O1gal-C4glc) and psi (C1gal-O1gal-C4glc-C5glc), which have values of -93.52 (13) and -144.83 (11) degrees, respectively, where the ring atom numbering conforms to the convention in which C1 is the anomeric C atom and C6 is the exocyclic hydroxymethyl (-CH2OH) C atom in both residues. The linkage geometry is similar to that observed in methyl beta-lactoside methanol solvate, (II), in which varphi is -88.4 (4) degrees and psi is -161.3 (4) degrees. As in (II), an intermolecular O3glc-H...O5gal hydrogen bond is observed in (I). The hydroxymethyl group conformation in both residues is gauche-trans, with torsion angles omegagal (O5gal-C5gal-C6gal-O6gal) and omega(glc) (O5glc-C5glc-C6glc-O6glc) of 69.15 (13) and 72.55 (14) degrees, respectively. The latter torsion angle differs substantially from that found for (II) [-54.6 (2) degrees; gauche-gauche]. Cocrystallization of methanol, which is hydrogen bonded to O6glc in the crystal structure of (II), presumably affects the hydroxymethyl conformation in the Glc residue in (II).


Acta Crystallographica Section C-crystal Structure Communications | 2005

Methyl 4-O-β-d-galacto­pyran­osyl α-d-gluco­pyran­oside (methyl α-lactoside)

Qingfeng Pan; Bruce C. Noll; Anthony S. Serianni

Methyl alpha-lactoside, C13H24O11, (I), is described by glycosidic torsion angles varphi (O5gal-C1gal-O1gal-C4glc) and psi (C1gal-O1gal-C4glc-C5glc), which have values of -93.52 (13) and -144.83 (11) degrees, respectively, where the ring atom numbering conforms to the convention in which C1 is the anomeric C atom and C6 is the exocyclic hydroxymethyl (-CH2OH) C atom in both residues. The linkage geometry is similar to that observed in methyl beta-lactoside methanol solvate, (II), in which varphi is -88.4 (4) degrees and psi is -161.3 (4) degrees. As in (II), an intermolecular O3glc-H...O5gal hydrogen bond is observed in (I). The hydroxymethyl group conformation in both residues is gauche-trans, with torsion angles omegagal (O5gal-C5gal-C6gal-O6gal) and omega(glc) (O5glc-C5glc-C6glc-O6glc) of 69.15 (13) and 72.55 (14) degrees, respectively. The latter torsion angle differs substantially from that found for (II) [-54.6 (2) degrees; gauche-gauche]. Cocrystallization of methanol, which is hydrogen bonded to O6glc in the crystal structure of (II), presumably affects the hydroxymethyl conformation in the Glc residue in (II).


Journal of Organic Chemistry | 2007

DFT and NMR studies of 2JCOH, 3JHCOH, and 3JCCOH Spin-Couplings in Saccharides : C-O torsional bias and H-bonding in aqueous solution

Hongqiu Zhao; Qingfeng Pan; Wenhui Zhang; Ian Carmichael; Anthony S. Serianni

Collaboration


Dive into the Qingfeng Pan's collaboration.

Top Co-Authors

Avatar
Top Co-Authors

Avatar

Bruce C. Noll

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar

Ian Carmichael

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar

Wenhui Zhang

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar

Xiaosong Hu

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

John G. Duman

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

Toby Turney

University of Notre Dame

View shared research outputs
Top Co-Authors

Avatar
Researchain Logo
Decentralizing Knowledge