Elucidation of the Interaction Mechanism between Organic Chiral Cages with Biomolecules through Nuclear Magnetic Resonance and Theoretical StudiesClick to copy article linkArticle link copied!
- Sara Sáez-FerreSara Sáez-FerreInstituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas (UPV-CSIC), Avda. de los Naranjos s/n, 46022 Valencia, SpainMore by Sara Sáez-Ferre
- Mercedes BoronatMercedes BoronatInstituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas (UPV-CSIC), Avda. de los Naranjos s/n, 46022 Valencia, SpainMore by Mercedes Boronat
- Ángel CantínÁngel CantínInstituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas (UPV-CSIC), Avda. de los Naranjos s/n, 46022 Valencia, SpainMore by Ángel Cantín
- Fernando Rey*Fernando Rey*E-mail: [email protected] (F.R.).Instituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas (UPV-CSIC), Avda. de los Naranjos s/n, 46022 Valencia, SpainMore by Fernando Rey
- Pascual Oña-Burgos*Pascual Oña-Burgos*E-mail: [email protected] (P.O.-B.).Instituto de Tecnología Química, Universitat Politècnica de València-Consejo Superior de Investigaciones Científicas (UPV-CSIC), Avda. de los Naranjos s/n, 46022 Valencia, SpainMore by Pascual Oña-Burgos
Abstract
A multinuclear NMR has been carried out to elucidate the mechanism of action of CC3-R box-type chiral materials for the separation of enantiomers, supported by theoretical calculations. The potential of these materials to be used as chiral resolution agents through NMR is evidence in this study.
Introduction
Experimental Section
General Considerations
Solution NMR Spectroscopy
![](/cms/10.1021/acs.jpcc.8b05069/asset/images/medium/jp-2018-05069u_m001.gif)
![](/cms/10.1021/acs.jpcc.8b05069/asset/images/medium/jp-2018-05069u_m002.gif)
X-ray Diffraction
Other Techniques
Density Functional Theory Studies
![](/cms/10.1021/acs.jpcc.8b05069/asset/images/medium/jp-2018-05069u_m003.gif)
Results
Experimental error in the D values was (±2)%.
The viscosity (η) used in the Stokes–Einstein equation was 0.410 × 10–3 kg m–1 s–1. Values of η were taken from http://www.knovel.com.
It is an estimation, deduced from the X-ray structure described in the ref (8).
NMR Spectroscopy Study of Host–Guest Interactions
Figure 1
Figure 1. Illustration of the selected guests (L/S red and D/R blue).
Figure 2
Figure 2. Illustration of the three selected signals of CC3-R.
δ 1H NMR [ppm] | |||||
---|---|---|---|---|---|
entry | [guest] [mM] | H1 | H2 | H3 | Δδ H1/H2/H3 |
1 | 8.165 | 7.891 | 3.38 | ||
2 | 1 mM l-1 | 8.165 | 7.891 | 3.381 | 0.000/0.000/0.001 |
3 | 3 mM l-1 | 8.168 | 7.907 | 3.397 | 0.003/0.016/0.017 |
4 | 10 mM l-1 | 8.176 | 7.935 | 3.428 | 0.011/0.044/0.048 |
5 | 1 mM d-1 | 8.165 | 7.893 | 3.382 | 0.000/0.002/0.002 |
6 | 3 mM d-1 | 8.166 | 7.900 | 3.390 | 0.001/0.011/0.010 |
7 | 10 mM d-1 | 8.173 | 7.926 | 3.416 | 0.008/0.035/0.036 |
8 | 1 mM l-2 | 8.167 | 7.894 | 3.385 | 0.002/0.003/0.005 |
9 | 3 mM l-2 | 8.168 | 7.899 | 3.391 | 0.003/0.008/0.011 |
10 | 10 mM l-2 | 8.167 | 7.904 | 3.400 | 0.002/0.013/0.020 |
11 | 1 mM d-2 | 8.165 | 7.890 | 3.379 | 0.000/0.001/0.001 |
12 | 3 mM d-2 | 8.165 | 7.895 | 3.385 | 0.000/0.004/0.005 |
13 | 10 mM d-2 | 8.165 | 7.902 | 3.396 | 0.000/0.011/0.016 |
14 | 1 mM l-3 | 8.165 | 7.890 | 3.379 | 0.000/–0.001/0.001 |
15 | 3 mM l-3 | 8.165 | 7.896 | 3.386 | 0.000/0.005/0.006 |
16 | 10 mM l-3 | 8.171 | 7.918 | 3.409 | 0.006/0.027/0.029 |
17 | 1 mM d-3 | 8.165 | 7.891 | 3.381 | 0.000/0.000/0.001 |
18 | 3 mM d-3 | 8.166 | 7.894 | 3.385 | 0.001/0.003/0.005 |
19 | 10 mM d-3 | 8.168 | 7.910 | 3.400 | 0.003/0.019/0.020 |
20 | 1 mM l-4 | 8.167 | 7.894 | 3.384 | 0.002/0.003/0.004 |
21 | 3 mM l-4 | 8.167 | 7.897 | 3.389 | 0.002/0.006/0.009 |
22 | 10 mM l-4 | 8.167 | 7.899 | 3.392 | 0.002/0.008/0.012 |
23 | 1 mM d-4 | 8.166 | 7.891 | 3.380 | 0.001/0.000/0.000 |
24 | 3 mM d-4 | 8.167 | 7.895 | 3.386 | 0.002/0.004/0.004 |
25 | 10 mM d-4 | 8.165 | 7.897 | 3.388 | 0.000/0.006/0.008 |
26 | 1 mM l-5 | 8.165 | 7.890 | 3.379 | 0.000/–0.001/0.001 |
27 | 3 mM l-5 | 8.165 | 7.891 | 3.380 | 0.000/0.000/0.000 |
28 | 10 mM l-5 | 8.166 | 7.894 | 3.386 | 0.001/0.003/0.006 |
29 | 1 mM d-5 | 8.166 | 7.891 | 3.380 | 0.001/0.000/0.000 |
30 | 3 mM d-5 | 8.166 | 7.892 | 3.382 | 0.001/0.001/0.002 |
31 | 10 mM d-5 | 8.165 | 7.893 | 3.385 | 0.000/0.002/0.005 |
32 | 10 mM S-6 | 8.162 | 7.889 | 3.378 | 0.003/0.003/0.002 |
33 | 10 mM R-6 | 8.162 | 7.889 | 3.377 | 0.003/0.003/0.003 |
Figure 3
Figure 3. 1H NMR data for H2 signal of the host with both enantiomers of ligand 1 at several concentrations.
δ 13C NMR [ppm] | δ 15N NMR [ppm] | ||||
---|---|---|---|---|---|
entry | [guest] [mM] | C1 | C2 | N3 | Δδ C1/C2/N3 |
1 | 158.65 | 129.03 | 341.3 | ||
2 | 3 mM l-1 | 159.10 | 129.20 | 338.7 | 0.45/0.17/2.60 |
3 | 10 mM l-1 | 159.10 | 129.40 | 337.3 | 0.45/0.37/4.00 |
4 | 3 mM d-1 | 159.00 | 129.13 | 339.4 | 0.35/0.10/1.90 |
5 | 10 mM d-1 | 159.18 | 129.48 | 337.6 | 0.53/0.45/3.70 |
6 | 3 mM l-2 | 158.8 | 129.20 | 339.5 | 0.15/0.17/1.80 |
7 | 10 mM l-2 | 158.9 | 129.30 | 338.6 | 0.25/0.27/2.70 |
8 | 3 mM d-2 | 158.65 | 129.09 | 340.0 | 0.00/0.06/1.30 |
9 | 10 mM d-2 | 158.94 | 129.33 | 339.3 | 0.29/0.30/2.00 |
10 | 3 mM l-3 | 158.80 | 129.15 | 339.8 | 0.15/0.12/1.50 |
11 | 10 mM l-3 | 159.00 | 129.30 | 338.2 | 0.35/0.27/3.10 |
12 | 3 mM d-3 | 158.78 | 129.11 | 340.4 | 0.13/0.08/0.90 |
13 | 10 mM d-3 | 159.80 | 129.13 | 339.1 | 0.15/0.10/2.20 |
14 | 3 mM l-4 | 158.94 | 129.13 | 340.9 | 0.29/0.10/0.40 |
25 | 10 mM l-4 | 159.00 | 129.32 | 339.6 | 0.34/0.29/1.70 |
16 | 3 mM d-4 | 158.86 | 129.06 | 341.2 | 0.21/0.03/0.10 |
17 | 10 mM d-4 | 158.96 | 129.27 | 340.3 | 0.31/0.24/1.00 |
18 | 3 mM l-5 | 158.65 | 129.09 | 341.0 | 0.00/0.06/0.30 |
19 | 10 mM l-5 | 158.80 | 129.20 | 340.4 | 0.15/0.17/0.90 |
20 | 3 mM d-5 | 158.74 | 129.03 | 341.6 | 0.09/0.00/0.30 |
21 | 10 mM d-5 | 158.80 | 129.11 | 340.6 | 0.15/0.08/0.70 |
22 | 10 mM S-6 | 158.69 | 129.06 | 341.0 | 0.04/0.03/0.30 |
23 | 10 mM R-6 | 158.67 | 129.04 | 341.1 | 0.02/0.01/0.20 |
δ 1H NMR (ppm) | |||||
---|---|---|---|---|---|
entry | [guest] [mM] | CC3-R mM | H1 | H2 | Δδ H1/H2 |
1 | l-1 | 4.388 | 1.488 | ||
2 | 3 mM l-1 | 2.5 | 4.3512 | 1.4863 | 0.0365/0.0017 |
3 | 10 mM l-1 | 2.5 | 4.311 | 1.4428 | 0.077/0.0452 |
4 | 3 mM d-1 | 2.5 | 4.3622 | 1.4865 | 0.0258/0.0015 |
5 | 10 mM d-1 | 2.5 | 4.351 | 1.4718 | 0.037/0.0165 |
6 | l-2 | 4.2914 | 1.44 | ||
7 | 3 mM l-2 | 2.5 | 4.2861 | 1.4413 | 0.0053/–0.0013 |
8 | 10 mM l-2 | 2.5 | 4.2808 | 1.4474 | 0.0106/–0.074 |
9 | 3 mM d-2 | 2.5 | 4.2905 | 1.4420 | 0.0009/–0.0020 |
10 | 10 mM d-2 | 2.5 | 4.2866 | 1.4415 | 0.0048/–0.0015 |
11 | l-3 | 4.2086 | 1.0246/0.96 | ||
12 | 3 mM l-3 | 2.5 | 4.208 | 1.0375/0.9709 | 0.0006/–0.0129/–0.0109 |
13 | 10 mM l-3 | 2.5 | 4.1815 | 1.0235/0.9604 | 0.0271/0.0011/–0.0004 |
14 | 3 mM d-3 | 2.5 | 4.2080 | 1.0369/0.9649 | 0.0006/–0.0123/–0.0049 |
15 | 10 mM d-3 | 2.5 | 4.1765 | 1.0148/0.9461 | 0.0321/0.0098/0.0139 |
16 | l-4 | 4.264 | 0.9836/0.9757 | ||
17 | 3 mM l-4 | 2.5 | 4.278 | 1.004/0.9917 | –0.014/–0.0204/–0.016 |
18 | 10 mM l-4 | 2.5 | 4.278 | 1.0098/0.9966 | –0.014/–0.0262/–0.0209 |
19 | 3 mM d-4 | 2.5 | 4.278 | 0.9962 | –0.014/–0.0126/–0.0205 |
20 | 10 mM d-4 | 2.5 | 4.278 | 1.0036 | –0.014/–0.02/–0.0279 |
21 | l-5 | 4.352 | |||
22 | 10 mM l-5 | 2.5 | 4.344 | 0.008 | |
23 | 10 mM d-5 | 2.5 | 4.350 | 0.002 | |
24 | 10 mM S-6 | 4.902 | 1.490 | ||
25 | 10 mM S-6 | 2.5 | 4.911 | 1.491 | –0.007/–0.001 |
26 | 10 mM R-6 | 2.5 | 4.911 | 1.490 | –0.007/0.000 |
Figure 4
Figure 4. 1H NMR region for the corresponding methyl group signal of ligand 1 at 3 mM and the host at 2.5 mM; (a) rac-1; (b) S-1; (c) R-1; and (d) isolated S-1.
entry | [guest] [3 mM] | COF [2.5 mM] | D [10–10 m2 s–1] | rH [Å] | guest free/guest inside |
---|---|---|---|---|---|
1 | X | 6.6 | 8.1 | ||
2 | l-1 | 22.3 | 2.4 | ||
3 | l-1 | X | 13.5 | 3.9 | 51.6/48.4 |
4 | d-1 | X | 14.2 | 3.7 | 56.1/43.9 |
5 | l-2 | 14 | 3.8 | ||
6 | l-2 | X | 12.4 | 4.3 | 78.2/21.8 |
7 | d-2 | X | 12.8 | 4.2 | 83.8/16.2 |
8 | l-3 | 13.1 | 4.1 | ||
9 | l-3 | X | 12.3 | 4.3 | 87.7/12.3 |
10 | d-3 | X | 12.8 | 4.2 | 91.4/8.6 |
11 | l-4 | 13.1 | 4.1 | ||
12 | l-4 | X | 12.2 | 4.4 | 86.2/13.8 |
13 | d-4 | X | 12.4 | 4.3 | 89.2/10.8 |
14 | l-5 | 15 | 3.5 | ||
15 | l-5 | X | 14.5 | 3.7 | 94/6 |
16 | d-5 | X | 14.9 | 3.6 | 98.8/1.2 |
17 | S-6 | 22.3 | 2.4 | ||
18 | S-6 | X | 21.1 | 2.5 | 95.5/4.5 |
19 | R-6 | X | 21.6 | 2.5 | 94.9/5.1 |
DFT Study of Host–Guest Interactions
Figure 5
Figure 5. DFT structures for the interaction CC3-R–guest; l-1 and d-1; l-2 and d-2.
entry | guest | Eint [kJ/mol] | rH–H1 [Å] | rH–H2 [Å] | rH–H3 [Å] |
---|---|---|---|---|---|
1 | l-1a | –76.9 | 2.519 | 3.540 | 2.838 |
2 | l-1b | –61.6 | 3.933 | 4.191 | 4.669 |
3 | d-1a | –74.8 | 2.923 | 3.633 | 3.414 |
4 | d-1b | –69.3 | 2.464 | 3.764 | 2.671 |
5 | l-2a | –78.9 | 2.548 | 3.451 | 2.837 |
6 | l-2b | –73.0 | 2.450 | 3.549 | 2.836 |
7 | d-2a | –72.7 | 2.265 | 3.631 | 2.512 |
8 | d-2b | –63.7 | 2.552 | 3.788 | 2.654 |
Conclusion
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.8b05069.
Experimental procedures; spectral and computational details; and cartesian coordinates for all optimized structures (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
Program Severo Ochoa SEV-2016-0683 is gratefully acknowledged. S.S-F. thanks MEC for his Severo Ochoa Grant SPV-2013-067884, P.O.-B. thanks MEC for his Ramón y Cajal contract RYC-2014-16620. M.B. and F.R. thank the financial support by the Spanish Government (MAT2017-82288-C2-1-P and MAT2015-71842-P). The authors thank the MULTY2HYCAT (EU-Horizon 2020 funded project under grant agreement no. 720783). The Electron Microscopy Service of the UPV is acknowledged for their help in sample characterization.
References
This article references 54 other publications.
- 1Pu, L. Enantioselective fluorescent sensors: a tale of BINOL. Acc. Chem. Res. 2012, 45, 150– 163, DOI: 10.1021/ar200048dGoogle Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtVWhu7%252FJ&md5=47d0d8ac638b37ea88aaaa8b173a7db8Enantioselective Fluorescent Sensors: A Tale of BINOLPu, LinAccounts of Chemical Research (2012), 45 (2), 150-163CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. The development of automated, high-throughput org. synthesis and screening techniques has created an urgent demand for methods that rapidly det. the enantiomeric compn. of chiral compds. Enantioselective fluorescent sensors offer the potential for real-time, high-sensitivity techniques for detg. enantiomeric data in high-throughput chiral assays. In this Account, the authors describe a range of fluorescent sensors derived from 1,1'-bi-2-naphthol (BINOL), a readily available biaryl compd. with axial chirality. BINOL can be used to construct structurally diverse, chiral fluorescent sensors to carry out highly enantioselective, sensitive recognition of chiral amino alcs., α-hydroxycarboxylic acids, and amino acid derivs. For example, the authors prepd. an (S)-BINOL deriv. whose 3,3'-positions are attached to two chiral amino alc. units, each having two Ph substituents. This compd. shows a fluorescence enhancement of 950-fold in the presence of (S)-mandelic acid but very little change in the presence of (R)-mandelic acid. It also allows the enantiomers of this α-hydroxycarboxylic acid to be visually discriminated by an enantioselective pptn. process. A structurally similar (S)-BINOL-amino alc. mol., but with three rather than two Ph substituents in each of the two amino alc. units, was found to exhibit generally enantioselective fluorescence responses toward structurally diverse α-hydroxycarboxylic acids. The authors further prepd. a pseudoenantiomeric analog of this compd. from (R)-H8BINOL, which has the opposite chiral configuration at both the biaryl center as well as the pendant amino alcs. These two compds. have opposite enantioselectivity in the recognition of a chiral substrate, with distinctly different fluorescence emission wavelengths. By mixing them together, the authors developed a pseudoenantiomeric sensor pair to facilitate chiral assays. Using this pseudoenantiomeric sensor pair allows both the concn. and the enantiomeric compn. of a substrate to be detd. in a single fluorescence measurement. The authors synthesized another compd. by ligating a terpyridine unit to BINOL and found that coordination of a Cu(II) ion to the terpyridine unit almost completely quenched its fluorescence. Displacement of the Cu2+ ion from this complex by chiral amino alcs. leads to enantioselective fluorescence enhancement. This BINOL-terpyridine-Cu(II) complex also exhibits enantioselective gel collapsing in the presence of chiral amino alcs., providing a new visual chiral discrimination method. When light-absorbing conjugated units are attached to the BINOL structure, the resulting multiarmed dendritic mols. show greatly amplified fluorescence responses. Thus, the light harvesting effect of dendrimers can be used to greatly increase the sensitivity of the fluorescent sensors. The progress described here demonstrates that highly enantioselective and sensitive fluorescent sensors can be obtained through a systematic study of the structure-property relation between the sensors and the substrates. These sensors show great potential for the development of rapid assays of chiral org. compds.
- 2Zhang, X.; Yin, J.; Yoon, J. Recent advances in development of chiral fluorescent and colorimetric sensors. Chem. Rev. 2014, 114, 4918– 4959, DOI: 10.1021/cr400568bGoogle Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXitVars70%253D&md5=00aec7910eb42651586cabcc683638e8Recent Advances in Development of Chiral Fluorescent and Colorimetric SensorsZhang, Xin; Yin, Jun; Yoon, JuyoungChemical Reviews (Washington, DC, United States) (2014), 114 (9), 4918-4959CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. In this review,the results of these efforts are organized based on the constitutional components and structural features of the sensors including small org. mol. based chiral chemosensors, metal complex based chiral probes, polymer based chiral probes, and nanomaterial based chiral sensors. Attention is also given to the chem. mechanisms involved in the chiral recognition and binding interactions involved in each sensing process.
- 3Simancas, R.; Dari, D.; Velamazan, N.; Navarro, M. T.; Cantin, A.; Jorda, J. L.; Sastre, G.; Corma, A.; Rey, F. Modular organic structure-directing agents for the synthesis of zeolites. Science 2010, 330, 1219– 1222, DOI: 10.1126/science.1196240Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVGrsbfP&md5=2b7a5738846b8dd1930fa1aa133f102fModular Organic Structure-Directing Agents for the Synthesis of ZeolitesSimancas, Raquel; Dari, Djamal; Velamazan, Noemi; Navarro, Maria T.; Cantin, Angel; Jorda, Jose L.; Sastre, German; Corma, Avelino; Rey, FernandoScience (Washington, DC, United States) (2010), 330 (6008), 1219-1222CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Org. structure-directing agents (OSDAs) were used to guide the formation of particular types of pores and channels during the synthesis of zeolites. The authors report that the use of highly versatile OSDAs based on phosphazenes has been successfully introduced for the synthesis of zeolites. This approach has made possible the synthesis of the elusive boggsite zeolite, which is formed by 10- and 12-ring intersecting channels. This topol. and these pore dimensions present interesting opportunities for catalysis in reactions of industrial relevance.
- 4Moliner, M.; Martínez, C.; Corma, A. Synthesis strategies for preparing useful small pore zeolites and zeotypes for gas separations and catalysis. Chem. Mater. 2014, 26, 246– 258, DOI: 10.1021/cm4015095Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXpsVOnuro%253D&md5=18314dab0b5f71260d7ebb504861bbadSynthesis Strategies for Preparing Useful Small Pore Zeolites and Zeotypes for Gas Separations and CatalysisMoliner, Manuel; Martinez, Cristina; Corma, AvelinoChemistry of Materials (2014), 26 (1), 246-258CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)A review; in the last years, the prepn. of small pore zeolites, esp. those presenting large cavities in their structure, have received much attention because they have shown excellent application in catalysis (such as methanol-to-olefins or selective catalytic redn. of NOx) and gas sepns. In the present review, we will focus on the diverse synthetic routes followed to direct the crystn. of small pore zeolites and, on the other hand, the most outstanding applications of those small pore microporous materials.
- 5Moliner, M.; Martínez, C.; Corma, A. Multipore zeolites: synthesis and catalytic applications. Angew. Chem., Int. Ed. 2015, 54, 3560– 3579, DOI: 10.1002/anie.201406344Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXktVKhu7c%253D&md5=07e31e0d3ff882898b85cdf5918fd327Multipore Zeolites: Synthesis and Catalytic ApplicationsMoliner, Manuel; Martinez, Cristina; Corma, AvelinoAngewandte Chemie, International Edition (2015), 54 (12), 3560-3579CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review; in the last few years, important efforts have been made to synthesize so-called "multipore" zeolites, which contain channels of different dimensions within the same cryst. structure. This is a very attractive subject, since the presence of pores of different sizes would favor the preferential diffusion of reactants and products through those different channel systems, allowing unique catalytic activities for specific chem. processes. In this Review we describe the most attractive achievements in the rational synthesis of multipore zeolites, contg. small to extra-large pores, and the improvements reported for relevant chem. processes when these multipore zeolites have been used as catalysts.
- 6Corma, A.; García, H.; Llabrés i Xamena, F. X. Engineering metal organic frameworks for heterogeneous catalysis. Chem. Rev. 2010, 110, 4606– 4655, DOI: 10.1021/cr9003924Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXktFers7Y%253D&md5=ce714f3d4475f4ac2a3d81b13931ccc3Engineering Metal Organic Frameworks for Heterogeneous CatalysisCorma, A.; Garcia, H.; Llabres i Xamena, F. X.Chemical Reviews (Washington, DC, United States) (2010), 110 (8), 4606-4655CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review, including the design of MOFs for catalysis and evaluation of its potential as catalyst as well as catalysis by MOFs with active metal sites, with reactive functional groups, and MOFs as host matrixes or nanometric reaction cavities.
- 7Dhakshinamoorthy, A.; Opanasenko, M.; Čejka, J.; Garcia, H. Metal organic frameworks as heterogeneous catalysts for the production of fine chemicals. Catal. Sci. Technol. 2013, 3, 2509– 2540, DOI: 10.1039/c3cy00350gGoogle Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsVSltrzL&md5=a0032e100abaa9a46a7ae87e3c7e2005Metal organic frameworks as heterogeneous catalysts for the production of fine chemicalsDhakshinamoorthy, Amarajothi; Opanasenko, Maksym; Cejka, Jiri; Garcia, HermenegildoCatalysis Science & Technology (2013), 3 (10), 2509-2540CODEN: CSTAGD; ISSN:2044-4753. (Royal Society of Chemistry)A review; this review focuses on the use of metal org. frameworks (MOFs) as catalysts for the synthesis of fine chems. While petrochem. is characterized by gas phase reactions, in which MOFs cannot compete with robust zeolites, MOFs are better suited for liq. phase reactions performed at moderate temps. These are the conditions typically employed for the prodn. of fine chems. characterized by being more complex and diverse mols. of low volatility, but with high added value. For the prepn. of this type of compd., MOFs offer the advantage of wide open porosity in the nanometer scale and a large void vol. In the present review we have summarized the reports that appeared up to early 2013 on the use of MOFs as catalysts in the liq. phase for the prodn. of fine chems., primarily classified according to the type of active site and the functional group formed in the reaction. Prospects for future development in this field are provided in the last section.
- 8Valvekens, P.; Vermoortele, F.; De Vos, D. Metal-organic frameworks as catalysts: the role of metal active sites. Catal. Sci. Technol. 2013, 3, 1435– 1445, DOI: 10.1039/c3cy20813cGoogle Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXns1KjsL4%253D&md5=38d269d004b12fe72deddab664c2c606Metal-organic frameworks as catalysts: the role of metal active sitesValvekens, Pieterjan; Vermoortele, Frederik; De Vos, DirkCatalysis Science & Technology (2013), 3 (6), 1435-1445CODEN: CSTAGD; ISSN:2044-4753. (Royal Society of Chemistry)A review. In this perspective first the use of metal-org. frameworks (MOFs) as supports for catalytic functions were critically compared with the possibilities offered by other classes of porous materials. Then the incidental or deliberate formation of active sites in MOF lattices were discuss and some strategies to control the no. and activity of these sites, ultimately resulting in MOF catalysts with improved performance, are reviewed.
- 9Jiang, J.; Yaghi, O. M. Brønsted Acidity in Metal-Organic Frameworks. Chem. Rev. 2015, 115, 6966– 6997, DOI: 10.1021/acs.chemrev.5b00221Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtVKisLvP&md5=c07defc470920b029260a98b76f241e0Bronsted Acidity in Metal-Organic FrameworksJiang, Juncong; Yaghi, Omar M.Chemical Reviews (Washington, DC, United States) (2015), 115 (14), 6966-6997CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review; Bronsted acidity in metal-org. frameworks is discussed.
- 10Seoane, B.; Castellanos, S.; Dikhtiarenko, A.; Kapteijn, F.; Gascon, J. Multi-scale crystal engineering of metal organic frameworks. Coord. Chem. Rev. 2016, 307, 147– 187, DOI: 10.1016/j.ccr.2015.06.008Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFWlsbbJ&md5=cf82259d468159c7347401a263b435b5Multi-scale crystal engineering of metal organic frameworksSeoane, Beatriz; Castellanos, Sonia; Dikhtiarenko, Alla; Kapteijn, Freek; Gascon, JorgeCoordination Chemistry Reviews (2016), 307 (Part_2), 147-187CODEN: CCHRAM; ISSN:0010-8545. (Elsevier B.V.)A review. During the last decade the no. of publications related to the synthesis of new metal-org. frameworks or coordination polymers has grown almost exponentially. Many of them are devoted to the study of the correlation between the mol. components (linkers and metal nodes or clusters) and the final properties of the resulting structure. Withal, the field of metal-org. frameworks has also witnessed important advances in the development of synthetic tools to control the particle size and shape and to obtain secondary porosity by applying knowledge from three disciplines: crystallog., coordination chem. and chem. engineering. These tools allow for crystal engineering beyond the mol. scale extending over the meso and macro scales, so that certain degree of multi-scale design is already possible. In this manner, MOFs' performance were improved in certain applications by choosing the optimal particle morphol. and dimensions that enhance the materials' properties and/or facilitate their implementation on functional devices. This review highlights the latest advances on MOF crystal engineering, with special emphasis on the meso and macro scales. After discussing some general considerations on the fundamentals of MOF crystn., the authors examine different synthetic approaches developed to tune the MOF particle size, shape and textural properties and the impact this multi-scale MOF crystal engineering showed so far in different applications. Finally, the authors' view on possible future research directions is outlined.
- 11Chen, L.; Luque, R.; Li, Y. Controllable design of tunable nanostructures inside metal-organic frameworks. Chem. Soc. Rev. 2017, 46, 4614– 4630, DOI: 10.1039/c6cs00537cGoogle Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXotVeqs7w%253D&md5=7a2d2dc717187c9fabbb83253103d6ceControllable design of tunable nanostructures inside metal-organic frameworksChen, Liyu; Luque, Rafael; Li, YingweiChemical Society Reviews (2017), 46 (15), 4614-4630CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)The controllable encapsulation of nanoentities (such as metal nanoparticles, quantum dots, polyoxometalates, org. and metallorg. mols., biomacromols., and metal-org. polyhedra) into metal-org. frameworks (MOFs) to form composite materials has attracted significant research interest in a variety of fields. These composite materials not only exhibit the properties of both the nanoentities and the MOFs but also display unique and synergistic functionalities. Tuning the sizes, compns., and shapes of nanoentities encapsulated in MOFs enables the final composites to exhibit superior performance to those of the sep. constituents for various applications. In this tutorial review article, we summarized the state-of-the-art development of MOFs contg. encapsulated tunable nanoentities, with special emphasis on the prepn. and synergistic properties of these composites.
- 12Zhang, Y.-B.; Su, J.; Furukawa, H.; Yun, Y.; Gándara, F.; Duong, A.; Zou, X.; Yaghi, O. M. Single-crystal structure of a covalent organic framework. J. Am. Chem. Soc. 2013, 135, 16336– 16339, DOI: 10.1021/ja409033pGoogle Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhs1Cktr3N&md5=ea22e0f58fcdda105e732e4fede57bc5Single-Crystal Structure of a Covalent Organic FrameworkZhang, Yue-Biao; Su, Jie; Furukawa, Hiroyasu; Yun, Yifeng; Gandara, Felipe; Duong, Adam; Zou, Xiaodong; Yaghi, Omar M.Journal of the American Chemical Society (2013), 135 (44), 16336-16339CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The crystal structure of a new covalent org. framework, termed COF-320, is detd. by single-crystal 3D electron diffraction using the rotation electron diffraction (RED) method for data collection. The COF crystals are prepd. by an imine condensation of tetra-(4-anilyl)-methane and 4,4'-biphenyldialdehyde in 1,4-dioxane at 120 °C to produce a highly porous 9-fold interwoven diamond net. COF-320 exhibits permanent porosity with a Langmuir surface area of 2400 m2/g and a methane total uptake of 15.0 wt. % (176 cm3/cm3) at 25 °C and 80 bar. The successful detn. of the structure of COF-320 directly from single-crystal samples is an important advance in the development of COF chem.
- 13Díaz, U.; Corma, A. Ordered covalent organic frameworks, COFs and PAFs. From preparation to application. Coord. Chem. Rev. 2016, 311, 85– 124, DOI: 10.1016/j.ccr.2015.12.010Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XjvFyntA%253D%253D&md5=209868f0abc6cdfc21da8c95b81ca7f5Ordered covalent organic frameworks, COFs and PAFs. From preparation to applicationDiaz, Urbano; Corma, AvelinoCoordination Chemistry Reviews (2016), 311 (), 85-124CODEN: CCHRAM; ISSN:0010-8545. (Elsevier B.V.)Covalent org. frameworks, COFs, and their derived sub-groups based on auto-assembly of exclusively arom. units, PAFs, are emerging into the advanced materials field due to their high free porous vol., structural regularity, robustness, hydrothermal stability, and functional variety. They present high gas uptake capacities and presence of stabilized active functions in the framework. This together with charged low-d. structures combined with their organization through π-conjugated system arrays, open the possibilities of COFs and PAFs to be used as effective materials for adsorption, selective sepn. and catalysis, and in nanotechnol. applications. This review will be focused on self-assembly synthesis mechanisms, physico-chem. characteristics, and applications of this class of promising covalent porous org. structures, out looking their possible future approaches and perspectives.
- 14Das, S.; Heasman, P.; Ben, T.; Qiu, S. Porous Organic Materials: Strategic Design and Structure-Function Correlation. Chem. Rev. 2017, 117, 1515– 1563, DOI: 10.1021/acs.chemrev.6b00439Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitFOhtLnP&md5=fe6aa70e2a25c48ac78974db574ff0f1Porous organic material strategic design and structure-function correlationDas, Saikat; Heasman, Patrick; Ben, Teng; Qiu, ShilunChemical Reviews (Washington, DC, United States) (2017), 117 (3), 1515-1563CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)Porous org. materials have garnered colossal interest with the scientific fraternity due to their excellent gas sorption performances, catalytic abilities, energy storage capacities, and other intriguing applications. This review encompasses the recent significant breakthroughs and the conventional functions and practices in the field of porous org. materials to find useful applications and imparts a comprehensive understanding of the strategic evolution of the design and synthetic approaches of porous org. materials with tunable characteristics. We present an exhaustive anal. of the design strategies with special emphasis on the topologies of cryst. and amorphous porous org. materials. In addn. to elucidating the structure-function correlation and state-of-the-art applications of porous org. materials, we address the challenges and restrictions that prevent us from realizing porous org. materials with tailored structures and properties for useful applications.
- 15Dawson, R.; Cooper, A. I.; Adams, D. J. Nanoporous organic polymer networks. Prog. Polym. Sci. 2012, 37, 530– 563, DOI: 10.1016/j.progpolymsci.2011.09.002Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtFeks7o%253D&md5=eeb7686ec5d0d748f906596074fdc8e3Nanoporous organic polymer networksDawson, Robert; Cooper, Andrew I.; Adams, Dave J.Progress in Polymer Science (2012), 37 (4), 530-563CODEN: PRPSB8; ISSN:0079-6700. (Elsevier Ltd.)A review. Nanoporous org. polymer networks are a class of materials consisting solely of the lighter elements in the periodic table. These materials have potential uses in areas such as storage, sepn., and catalysis. Here, we review the different classes of nanoporous polymer networks including covalent org. frameworks, hypercrosslinked polymers, conjugated microporous polymers, and polymers of intrinsic microporosity. The growing variety in synthetic routes to these materials allows a range of different polymer networks to be formed, including cryst. and amorphous structures. It is also possible to incorporate many different kinds of functional groups in a modular fashion. So far, most networks have been examd. from the perspective of gas sorption, and this area is discussed critically and in depth in this review. The use of nanoporous org. polymers for applications such as catalysis and sepns. is an important developing area, and we discuss recent developments as well as highlighting potential future opportunities.
- 16Tan, L.; Tan, B. Hypercrosslinked porous polymer materials: design, synthesis, and applications. Chem. Soc. Rev. 2017, 46, 3322– 3356, DOI: 10.1039/c6cs00851hGoogle Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXjtFelu7s%253D&md5=8a267d5ef3aa25a59cce8cfc98e19182Hypercrosslinked porous polymer materials: design, synthesis, and applicationsTan, Liangxiao; Tan, BienChemical Society Reviews (2017), 46 (11), 3322-3356CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Hypercrosslinked polymers (HCPs) are a series of permanent microporous polymer materials initially reported by Davankov, and have received an increasing level of research interest. In recent years, HCPs have experienced rapid growth due to their remarkable advantages such as diverse synthetic methods, easy functionalization, high surface area, low cost reagents and mild operating conditions. Judicious selection of monomers, appropriate length crosslinkers and optimized reaction conditions yielded a well-developed polymer framework with an adjusted porous topol. Post fabrication of the as developed network facilitates the incorporation of various chem. functionalities that may lead to interesting properties and enhance the selection toward a specific application. To date, numerous HCPs have been prepd. by post-crosslinking polystyrene-based precursors, one-step self-polycondensation or external crosslinking strategies. The advent of these methodologies has prompted researchers to construct well-defined porous polymer networks with customized micromorphol. and functionalities. In this review, we describe not only the basic synthetic principles and strategies of HCPs, but also the advancements in the structural and morphol. study as well as the frontiers of potential applications in energy and environmental fields such as gas storage, carbon capture, removal of pollutants, mol. sepn., catalysis, drug delivery, sensing, etc.
- 17Bojdys, M. J.; Hasell, T.; Severin, N.; Jelfs, K. E.; Rabe, J. P.; Cooper, A. I. Porous organic cage crystals: characterising the porous crystal surface. Chem. Commun. 2012, 48, 11948– 11950, DOI: 10.1039/c2cc36602aGoogle Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhslCgtb%252FP&md5=986f5e1ef87ea6017720282e6a2fae1cPorous organic cage crystals: characterizing the porous crystal surfaceBojdys, Michael J.; Hasell, Tom; Severin, Nikolai; Jelfs, Kim E.; Rabe, Juergen P.; Cooper, Andrew I.Chemical Communications (Cambridge, United Kingdom) (2012), 48 (98), 11948-11950CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The characterization of porous cryst. solids often relies on single crystal x-ray diffraction, which does not give direct information about the surface of the material. Here, crystals of a porous org. cage, CC3-R, were studied by at. force microscopy and shown to possess two distinct gas-solid interfaces, proving that the bulk crystal structure is preserved at the porous crystal surface.
- 18Tozawa, T.; Jones, J. T. A.; Swamy, S. I.; Jiang, S.; Adams, D. J.; Shakespeare, S.; Clowes, R.; Bradshaw, D.; Hasell, T.; Chong, S. Y.; Tang, C.; Thompson, S.; Parker, J.; Trewin, A.; Bacsa, J.; Slawin, A. M. Z.; Steiner, A.; Cooper, A. I. Porous organic cages. Nat. Mater. 2009, 8, 973– 978, DOI: 10.1038/nmat2545Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhsVGlt7vP&md5=34d8de0ea8a080d32ce9d3f8b6abc9c3Porous organic cagesTozawa, Tomokazu; Jones, James T. A.; Swamy, Shashikala I.; Jiang, Shan; Adams, Dave J.; Shakespeare, Stephen; Clowes, Rob; Bradshaw, Darren; Hasell, Tom; Chong, Samantha Y.; Tang, Chiu; Thompson, Stephen; Parker, Julia; Trewin, Abbie; Bacsa, John; Slawin, Alexandra M. Z.; Steiner, Alexander; Cooper, Andrew I.Nature Materials (2009), 8 (12), 973-978CODEN: NMAACR; ISSN:1476-1122. (Nature Publishing Group)Porous materials are important in a wide range of applications including mol. sepns. and catalysis. Covalently bonded org. cages can assemble into cryst. microporous materials. The porosity is prefabricated and intrinsic to the mol. cage structure, as opposed to being formed by noncovalent self-assembly of nonporous sub-units. The three-dimensional connectivity between the cage windows is controlled by varying the chem. functionality such that either nonporous or permanently porous assemblies can be produced. Surface areas and gas uptakes for the latter exceed comparable mol. solids. One of the cages can be converted by recrystn. to produce either porous or nonporous polymorphs with apparent Brunauer-Emmett-Teller surface areas of 550 and 23 m2 g-1, resp. These results suggest design principles for responsive porous org. solids and for the modular construction of extended materials from prefabricated mol. pores.
- 19Chen, L.; Reiss, P. S.; Chong, S. Y.; Holden, D.; Jelfs, K. E.; Hasell, T.; Little, M. A.; Kewley, A.; Briggs, M. E.; Stephenson, A.; Thomas, K. M.; Armstrong, J. A.; Bell, J.; Busto, J.; Noel, R.; Liu, J.; Strachan, D. M.; Thallapally, P. K.; Cooper, A. I. Separation of rare gases and chiral molecules by selective binding in porous organic cages. Nat. Mater. 2014, 13, 954– 960, DOI: 10.1038/nmat4035Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFGqu77K&md5=24786fc90209b72d87021b8626a2422fSeparation of rare gases and chiral molecules by selective binding in porous organic cagesChen, Linjiang; Reiss, Paul S.; Chong, Samantha Y.; Holden, Daniel; Jelfs, Kim E.; Hasell, Tom; Little, Marc A.; Kewley, Adam; Briggs, Michael E.; Stephenson, Andrew; Thomas, K. Mark; Armstrong, Jayne A.; Bell, Jon; Busto, Jose; Noel, Raymond; Liu, Jian; Strachan, Denis M.; Thallapally, Praveen K.; Cooper, Andrew I.Nature Materials (2014), 13 (10), 954-960CODEN: NMAACR; ISSN:1476-1122. (Nature Publishing Group)A porous org. cage mol. having unprecedented performance in the solid state for the sepn. of rare gases, such as krypton and xenon, is reported. The selectivity arises from a precise size match between the rare gas and the org. cage cavity, as predicted by mol. simulations. Breakthrough expts. demonstrate real practical potential for the sepn. of krypton, xenon and radon from air at concns. of only a few ppm. Selective binding of chiral org. mols. such as 1-phenylethanol is also demonstrated, suggesting applications in enantioselective sepn.
- 20Briggs, M. E.; Slater, A. G.; Lunt, N.; Jiang, S.; Little, M. A.; Greenaway, R. L.; Hasell, T.; Battilocchio, C.; Ley, S. V.; Cooper, A. I. Dynamic flow synthesis of porous organic cages. Chem. Commun. 2015, 51, 17390– 17393, DOI: 10.1039/c5cc07447aGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhs1eiurnE&md5=03eb47fc8c1bbe1cfcaf09745cc5cf6aDynamic flow synthesis of porous organic cagesBriggs, Michael E.; Slater, Anna G.; Lunt, Neil; Jiang, Shan; Little, Marc A.; Greenaway, Rebecca L.; Hasell, Tom; Battilocchio, Claudio; Ley, Steven V.; Cooper, Andrew I.Chemical Communications (Cambridge, United Kingdom) (2015), 51 (98), 17390-17393CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The dynamic covalent synthesis of two imine-based porous org. cages was successfully transferred from batch to continuous flow. The same flow reactor was then used to scramble the constituents of these two cages in differing ratios to form cage mixts. Preparative HPLC purifn. of one of these mixts. allowed rapid access to a desymmetrised cage mol.
- 21Jiang, S.; Jones, J. T. A.; Hasell, T.; Blythe, C. E.; Adams, D. J.; Trewin, A.; Cooper, A. I. Porous organic molecular solids by dynamic covalent scrambling. Nat. Commun. 2011, 2, 207, DOI: 10.1038/ncomms1207Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3M3htlSrtQ%253D%253D&md5=45e70ff90a4fbf3282cd9d8810d9f5e3Porous organic molecular solids by dynamic covalent scramblingJiang Shan; Jones James T A; Hasell Tom; Blythe Charlotte E; Adams Dave J; Trewin Abbie; Cooper Andrew INature communications (2011), 2 (), 207 ISSN:.The main strategy for constructing porous solids from discrete organic molecules is crystal engineering, which involves forming regular crystalline arrays. Here, we present a chemical approach for desymmetrizing organic cages by dynamic covalent scrambling reactions. This leads to molecules with a distribution of shapes which cannot pack effectively and, hence, do not crystallize, creating porosity in the amorphous solid. The porous properties can be fine tuned by varying the ratio of reagents in the scrambling reaction, and this allows the preparation of materials with high gas selectivities. The molecular engineering of porous amorphous solids complements crystal engineering strategies and may have advantages in some applications, for example, in the compatibilization of functionalities that do not readily cocrystallize.
- 22Little, M. A.; Chong, S. Y.; Schmidtmann, M.; Hasell, T.; Cooper, A. I. Guest control of structure in porous organic cages. Chem. Commun. 2014, 50, 9465– 9468, DOI: 10.1039/c4cc04158eGoogle Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFCgtb3E&md5=a039e2cc30ed71763c6f7d4f21eaa7edGuest control of structure in porous organic cagesLittle, Marc A.; Chong, Samantha Y.; Schmidtmann, Marc; Hasell, Tom; Cooper, Andrew I.Chemical Communications (Cambridge, United Kingdom) (2014), 50 (67), 9465-9468CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Two porous org. cages with different thermodn. polymorphs were induced by co-solvents to interchange their crystal packing modes, thus achieving guest-mediated control over solid-state porosity. In situ crystallog. allows the effect of the co-solvent guests on these structural interconversions to be understood.
- 23Kewley, A.; Stephenson, A.; Chen, L.; Briggs, M. E.; Hasell, T.; Cooper, A. I. Porous organic cages for gas chromatography separations. Chem. Mater. 2015, 27, 3207– 3210, DOI: 10.1021/acs.chemmater.5b01112Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXlvV2gsLc%253D&md5=885577c1c343019e8f7781f17243ff82Porous Organic Cages for Gas Chromatography SeparationsKewley, Adam; Stephenson, Andrew; Chen, Linjiang; Briggs, Michael E.; Hasell, Tom; Cooper, Andrew I.Chemistry of Materials (2015), 27 (9), 3207-3210CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)Chromatog. is an important process for characterizing and purifying org. mols., but some mixts. remain challenging to sep. The authors report the 1st use of a soln.-processable porous org. cage mol., (I, CC3), as a stationary phase for gas chromatog. Capillary columns were coated with CC3 using a simple, static coating method. These columns are versatile and can sep. alkanes, arom. mixts., and chiral mols. Atomistic simulations indicate that the efficient shape sorting of branched hexane isomers by CC3 results from a combination of kinetic diffusion and surface-interaction effects.
- 24Briggs, M. E.; Cooper, A. I. A perspective on the synthesis, purification, and characterization of porous organic cages. Chem. Mater. 2017, 29, 149– 157, DOI: 10.1021/acs.chemmater.6b02903Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhsVKrtrnI&md5=297fd2f519911113fc6e20d278c97074A Perspective on the Synthesis, Purification, and Characterization of Porous Organic CagesBriggs, Michael E.; Cooper, Andrew I.Chemistry of Materials (2017), 29 (1), 149-157CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)A review with authors data. Porous org. cages present many opportunities in functional materials chem., but the synthetic challenges for these mol. solids are somewhat different from those faced in the areas of metal-org. frameworks, covalent-org. frameworks, or porous polymer networks. Here, we highlight the practical methods that we have developed for the design, synthesis, and characterization of imine porous org. cages using CC1 and CC3 as examples. The key points are transferable to other cages, and this perspective should serve as a practical guide to researchers who are new to this field.
- 25Xie, S.-M.; Zhang, J.-H.; Fu, N.; Wang, B.-J.; Chen, L.; Yuan, L.-M. A chiral porous organic cage for molecular recognition using gas chromatography. Anal. Chim. Acta 2016, 903, 156– 163, DOI: 10.1016/j.aca.2015.11.030Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFantrfN&md5=9975aeabca89d90620731fac00a0f3dcA chiral porous organic cage for molecular recognition using gas chromatographyXie, Sheng-Ming; Zhang, Jun-Hui; Fu, Nan; Wang, Bang-Jin; Chen, Ling; Yuan, Li-MingAnalytica Chimica Acta (2016), 903 (), 156-163CODEN: ACACAM; ISSN:0003-2670. (Elsevier B.V.)Mol. org. cages as shape-persistent org. mols. with permanent and accessible cavities have attracted a lot of interest because of their importance as host-guest systems. Herein, the authors report a chiral porous org. cage (POC) CC9 dild. with a polysiloxane OV-1701 to fabricate a CC9-coated capillary column, which was used for the high-resoln. gas chromatog. sepn. of org. compds., including positional isomers and racemates. On the porous org. cage CC9-coated capillary column, a large no. of racemic compds. such as chiral alcs., esters, ethers and epoxides can be resolved without derivatization. By comparing the chiral recognition ability of the CC9-coated column with the com. available β-DEX 120 column and the POC CC3-R coated column recently reported by the authors' group, the CC9-coated column offered good resoln. during the sepn. of some racemates, that were not sepd. using the β-DEX 120 column or POC CC3-R coated column. Therefore, the CC9-coated column can be complementary to the β-DEX 120 column and CC3-R coated column. The CC9-coated column exhibited great potential for application in the sepn. of positional isomers and enantiomers with great selectivity, high resoln. and good reproducibility.
- 26Zhang, J.-H.; Xie, S.-M.; Chen, L.; Wang, B.-J.; He, P.-G.; Yuan, L.-M. Homochiral porous organic cage with high selectivity for the separation of racemates in gas chromatography. Anal. Chem. 2015, 87, 7817– 7824, DOI: 10.1021/acs.analchem.5b01512Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFWqtLjO&md5=aab5bfe5ea412eac7aeb48d04e3048edHomochiral Porous Organic Cage with High Selectivity for the Separation of Racemates in Gas ChromatographyZhang, Jun-Hui; Xie, Sheng-Ming; Chen, Ling; Wang, Bang-Jin; He, Pin-Gang; Yuan, Li-MingAnalytical Chemistry (Washington, DC, United States) (2015), 87 (15), 7817-7824CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)Porous org. mol. cages as a new type of porous materials have attracted a tremendous attention for their potential applications in recent years. Here the authors report the use of a homochiral porous org. cage (POC) (CC3-R) dild. with a polysiloxane (OV-1701) as a stationary phase for high-resoln. gas chromatog. (GC) with excellent enantioselectivity. A large no. of optical isomers were resolved without derivatization, including chiral alcs., diols, amines, alc. amines, esters, ketones, ethers, halohydrocarbons, org. acids, amino acid Me esters, and sulfoxides. Compared with com. β-DEX 120 and Chirasil-L-Val columns, the CC3-R coated capillary column offered more preeminent enantioselectivity. CC3-R also exhibits good selectivity for the sepn. of isomers, linear alkanes, alcs., and arom. hydrocarbons. The excellent resoln. ability, repeatability, and thermal stability make CC3-R a promising candidate as a novel stationary phase for GC. The study described herein first proves useful com. Also porous org. mol. materials will become more attractive in sepn. science.
- 27Stilbs, P. Fourier transform pulsed-gradient spin-echo studies of molecular diffusion. Prog. Nucl. Magn. Reson. Spectrosc. 1987, 19, 1– 45, DOI: 10.1016/0079-6565(87)80007-9Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXls1Shsg%253D%253D&md5=ef17c6f8b4d2f92752cae2ee3bbf82c1Fourier transform pulsed-gradient spin-echo studies of molecular diffusionStilbs, PeterProgress in Nuclear Magnetic Resonance Spectroscopy (1987), 19 (1), 1-45CODEN: PNMRAT; ISSN:0079-6565.A review with 193 refs. is given on NMR methods (esp. Fourier-transform, pulsed-field-gradient, spin-echo NMR) in measuring mol. diffusion (self-diffusion, mutual diffusion, and multicomponent diffusion) in various types of systems (including liqs., liq. mixts., polymer and polyelectrolyte solns., gels, surfactant aq. solns., and emulsions).
- 28Tyrrell, H. J. V.; Harris, R. K. Diffusion in Liquids; Butterworths: London, 1984.Google ScholarThere is no corresponding record for this reference.
- 29Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215– 241, DOI: 10.1007/s00214-007-0310-xGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 30Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650– 654, DOI: 10.1063/1.438955Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3cXpvFyitA%253D%253D&md5=eb331bad0222adcbe7cad51527273725Self-consistent molecular orbital methods. XX. A basis set for correlated wave functionsKrishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A.Journal of Chemical Physics (1980), 72 (1), 650-4CODEN: JCPSA6; ISSN:0021-9606.A contracted Gaussian basis set (6-311G**) is developed by optimizing exponents and coeffs. at the Moller-Plesset (MP) second-order level for the ground states of first-row atoms. This has a triple split in the valence s and p shells together with a single set of uncontracted polarization functions on each atom. The basis is tested by computing structures and energies for some simple mols. at various levels of MP theory and comparing with expt.
- 31Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford CT, 2009.Google ScholarThere is no corresponding record for this reference.
- 32Stejskal, E. O.; Tanner, J. E. Spin Diffusion Measurements: Spin Echoes in the Presence of a Time-Dependent Field Gradient. J. Chem. Phys. 1965, 42, 288– 292, DOI: 10.1063/1.1695690Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF2MXisFOrtQ%253D%253D&md5=c44b1de92c67b6fc3245225937c4ddd0Spin diffusion measurements: spin echoes in the presence of a time-dependent field gradientStejskal, E. O.; Tanner, J. E.Journal of Chemical Physics (1965), 42 (1), 288-92CODEN: JCPSA6; ISSN:0021-9606.A derivation is given of the effect of a time-dependent magnetic field gradient on the spin-echo expt., particularly in the presence of spin diffusion. There are several reasons for preferring certain kinds of time-dependent magnetic field gradients to the more usual steady gradient. If the gradient is reduced during the radio-frequency pulses, H1 need not be particularly large; if the gradient is small at the time of the echo, the echo will be broad and its amplitude easy to measure. Both of these relaxations of restrictions on the measurement of diffusion coeffs. by the spin-echo technique serve to extend its range of applicability. A pulsed gradient can be recommended when it is crit. to define the precise time period over which diffusion is being measured. The theoretical expression derived was verified exptl. for several choices of time-dependent magnetic field gradient. An app. is described suitable for the production of pulsed gradients with amplitudes as large as 100 gauss cm. The diffusion coeff. of dry glycerol at 26° ± 1° is (2.5 ± 0.2) × 10-3 cm.2/sec., a value smaller than can ordinarily be measured by the steady gradient method.
- 33Tanner, J. E. Use of the stimulated echo in NMR diffusion studies. J. Chem. Phys. 1970, 52, 2523– 2526, DOI: 10.1063/1.1673336Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXptlGgsw%253D%253D&md5=7390bc823359258bfb885733b25de844Use of the stimulated echo in NMR diffusion studiesTanner, John E.Journal of Chemical Physics (1970), 52 (5), 2523-6CODEN: JCPSA6; ISSN:0021-9606.The stimulated echo in a three-radio-frequency-pulse expt. is useful in extending the range of measurement of diffusion coeffs. to more viscous substances or the measurement of barrier sepns. to wider spacings in systems where the diffusing substance has T1 > T2. The spin-echo attenuation due to self-diffusion is derived for the general case of a time-dependent field gradient, and the result is found exptl. to be correct for the special case of a field gradient applied in 2 equal, square pulse.
- 34Johnson, C. S. Diffusion ordered nuclear magnetic resonance spectroscopy: principles and applications. Magn. Reson. Spectrosc. 1999, 34, 203– 256, DOI: 10.1016/s0079-6565(99)00003-5Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXivVOltbc%253D&md5=956048c606a7721394fc41687da69820Diffusion ordered nuclear magnetic resonance spectroscopy: principles and applicationsJohnson, C. S., Jr.Progress in Nuclear Magnetic Resonance Spectroscopy (1999), 34 (3,4), 203-256CODEN: PNMRAT; ISSN:0079-6565. (Elsevier Science B.V.)A review with 146 refs.
- 35Chen, A.; Wu, D.; Johnson, C. S. Determination of molecular weight distributions for polymers by diffusion-ordered NMR. J. Am. Chem. Soc. 1995, 117, 7965– 7970, DOI: 10.1021/ja00135a015Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmvVegsbc%253D&md5=ed023c8af0c849c7a607da6fc11a93edDetermination of Molecular Weight Distributions for Polymers by Diffusion-Ordered NMRChen, Aidi; Wu, Donghui; Johnson, Charles S., Jr.Journal of the American Chemical Society (1995), 117 (30), 7965-70CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Diffusion-ordered NMR spectroscopy, a technique based on pulsed field gradient NMR (PFGNMR), was used to characterize mol. wt. distributions for samples of poly(ethylene oxide) in D2O. The distribution of diffusion coeffs. G(D) was obtained by anal. of PFGNMR data with a modified version of the well-known constrained regularization program CONTIN. Regularization was considerably improved for broad distributions by setting the wts. cm in CONTIN's quadrature formula equal to (Dm/Dmax)xm, where Dmax is the diffusion coeff. corresponding to the max. in G(Dm)Dm and m is an index. Here xm varies linearly from +2 to -2 with log(Dm) across the distribution. This amts. to enhancing low amplitude regions of G(D)D during anal. The estd. distribution was then converted to the mass weighted distribution of mol. wts. by means of the relation D = 10-7.62 M-0.62 (with D in units of m2 s-1) obtained from expts. on monodisperse ref. stds. In this study spin relaxation rates were independent of mol. wts. and intermol. averaging effects were insignificant. As an illustration, mol. wt. distributions were detd. for two broadly distributed samples. The no. and wt. av. mol. wts. and the polydispersities agreed well with values provided by the manufacturer when the PFGNMR data sets had signal-to-noise ratios greater than 500.
- 36Avram, L.; Cohen, Y. Diffusion NMR of molecular cages and capsules. Chem. Soc. Rev. 2015, 44, 586– 602, DOI: 10.1039/c4cs00197dGoogle Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtlCjurzL&md5=8f40f901cfd7bac40cb022bda464359fDiffusion NMR of molecular cages and capsulesAvram, Liat; Cohen, YoramChemical Society Reviews (2015), 44 (2), 586-602CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. In the last decade diffusion NMR and diffusion ordered spectroscopy (DOSY) have become important anal. tools for the characterization of supramol. systems in soln. Diffusion NMR can be used to glean information on the (effective) size and shape of mol. species, as well as to probe inter-mol. interactions and can be used to est. the assocn. const. (Ka) of a complex. In addn., the diffusion coeff., as obtained from diffusion NMR, is a much more intuitive parameter than the chem. shift for probing self-assocn., aggregation and inter-mol. interactions. The diffusion coeff. may be an even more important anal. parameter in systems in which the formed supramol. entity has the same symmetry as its building units, when there is a large change in the mol. wt., where many mol. species are involved in the formation of the supramol. systems, and when proton transfer may occur which, in turn, may affect the chem. shift. Some of the self-assembled mol. capsules and cages prepd. in the last decade represent such supramol. systems and in the present review, following a short introduction on diffusion NMR, we survey the contribution of diffusion NMR and DOSY in the field of mol. containers and capsules. We will 1st focus on the role played by diffusion NMR in the field of H-bond driven self-assembled capsules. We then survey the contributions of diffusion NMR and DOSY to the study and characterization of metal-ligand cages and capsules. Finally, we describe a few recent applications of diffusion NMR in the field of hydrophobic, electrostatic and covalent containers.
- 37Åslund, I.; Nowacka, A.; Nilsson, M.; Topgaard, D. Filter-exchange PGSE NMR determination of cell membrane permeability. J. Magn. Reson. 2009, 200, 291– 295, DOI: 10.1016/j.jmr.2009.07.015Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BD1MnisVGnuw%253D%253D&md5=cf19b3e4a54adf6709dee41ae77a48dbFilter-exchange PGSE NMR determination of cell membrane permeabilityAslund Ingrid; Nowacka Agnieszka; Nilsson Markus; Topgaard DanielJournal of magnetic resonance (San Diego, Calif. : 1997) (2009), 200 (2), 291-5 ISSN:.A new PGSE NMR sequence is introduced for measuring diffusive transport across the plasma membrane of living cells. A "diffusion filter" and a variable mixing time precedes a standard PGSE block for diffusion encoding of the NMR signal. The filter is a PGSE block optimized for selectively removing the magnetization of the extracellular water. With increasing mixing time the intra- and extracellular components approach their equilibrium fractional populations. The rate of exchange can be measured using only a few minutes of instrument time. Water exchange over the plasma membrane of starved yeast cells is studied in the temperature range +5 to +32 degrees C.
- 38Occhipinti, P.; Griffiths, P. C. Quantifying diffusion in mucosal systems by pulsed-gradient spin-echo NMR. Adv. Drug Delivery Rev. 2008, 60, 1570– 1582, DOI: 10.1016/j.addr.2008.08.006Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhsVSjsL%252FM&md5=9d3a2f6cec6bfcac0575c7d0c61b1583Quantifying diffusion in mucosal systems by pulsed-gradient spin-echo NMROcchipinti, Paola; Griffiths, Peter C.Advanced Drug Delivery Reviews (2008), 60 (15), 1570-1582CODEN: ADDREP; ISSN:0169-409X. (Elsevier B.V.)A review. Mucus, a thick and slimy secretion produced by submucosal cells, covers many epithelial surfaces in mammalian organs and prevents foreign particles that enter the body from accessing cells. However, the mucus layer also represents a potential barrier to the efficient delivery of nano-sized drug delivery systems (polyplexes, lipoplexes, particles) to the underlying mucosal epithelium. Many studies have considered the ability of nano-sized particles and polymers to diffuse within the mucosal network using a range of different techniques, including multiple-particle tracking (MPT), diffusion chamber studies and fluorescence recovery after photobleaching (FRAP). This review highlights the current understanding of the interaction of the diffusion of nano-sized structures within mucosal networks. Moreover, this article presents an introduction to pulsed-gradient spin-echo NMR (PGSE-NMR), a potential new tool to investigate the mobility of mol. species through mucosal networks and related biol. gels.
- 39Marega, R.; Aroulmoji, V.; Dinon, F.; Vaccari, L.; Giordani, S.; Bianco, A.; Murano, E.; Prato, M. Diffusion-ordered NMR spectroscopy in the structural characterization of functionalized carbon nanotubes. J. Am. Chem. Soc. 2009, 131, 9086– 9093, DOI: 10.1021/ja902728wGoogle Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXmtlWhu7Y%253D&md5=2d08688be42573ccff4023551e4b38a3Diffusion-Ordered NMR Spectroscopy in the Structural Characterization of Functionalized Carbon NanotubesMarega, Riccardo; Aroulmoji, Vincent; Dinon, Francesca; Vaccari, Lisa; Giordani, Silvia; Bianco, Alberto; Murano, Erminio; Prato, MaurizioJournal of the American Chemical Society (2009), 131 (25), 9086-9093CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The emerging applications of functionalized C nanotubes (CNTs) in various research domains necessitate the use of many different anal. techniques to confirm their structural modifications in a fast and reliable manner. Thus far, NMR spectroscopy was not among the main tools for characterization of organically modified C nanostructures. 1H anal. is limited because the signals in these derivs. are typically weak and broad, resulting in uncertainties of a few ppm, and because of the strong interference of residual solvent signals. To overcome these limitations, the authors studied the applicability of 1H NMR spectroscopy based on gradient-edited diffusion pulse sequences (1-dimensional diffusion-ordered spectroscopy, DOSY) in the characterization of CNT derivs. In general, diffusion NMR expts. allow the sepn. of NMR signals of different species present in a mixt., according to their own diffusion coeffs., merging spectroscopy information with size anal. A selected set of CNT derivs. was synthesized and analyzed using 1-dimensional DOSY expts. by applying strong magnetic field gradients (up to 42.6 G cm-1). Colorimetric tests (i.e., Kaiser test) and TGA anal. support the NMR findings, which are related to isolated and/or bundled short SWNTs, from TEM and AFM characterization. The overall results show that the diffusion-based NMR spectroscopy is a fast and promising approach for the characterization of covalently modified CNT derivs.
- 40Canzi, G.; Mrse, A. A.; Kubiak, C. P. Diffusion-ordered NMR spectroscopy as a reliable alternative to TEM for determining the size of gold nanoparticles in organic solutions. J. Phys. Chem. C 2011, 115, 7972– 7978, DOI: 10.1021/jp2008557Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXktFelsbk%253D&md5=4c42a9f46d6d381a10a00381fe8c1ac3Diffusion-Ordered NMR Spectroscopy as a Reliable Alternative to TEM for Determining the Size of Gold Nanoparticles in Organic SolutionsCanzi, Gabriele; Mrse, Anthony A.; Kubiak, Clifford P.Journal of Physical Chemistry C (2011), 115 (16), 7972-7978CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Diffusion-ordered spectroscopy is used to det. gold nanoparticle sizes. Traditional characterization of nanoparticles has centered on imaging by electron microscopy and plasmon resonance absorption in UV-visible electronic spectra. The authors present a convenient method to characterize gold nanoparticles using diffusion-ordered NMR spectroscopy (DOSY). 2-dimensional DOSY NMR is used to calc. diffusion consts. and the diam. of solubilized gold nanoparticles capped with 1-dodecanethiol (C12) or 1-octanethiol (C8) in three deuterated solvents. The distributions of nanoparticle sizes strongly correlate with TEM image anal. C12 and C8 capped nanoparticle sizes are 4.6 and 2.7 nm by TEM as compared to ests. of 4.6 ± 0.3 and 2.5 ± 0.1 nm based on 2-dimensional DOSY NMR data. Reliable size characterization of nanoparticles with NMR active nuclei (1H in this study) in their protective groups (alkane thiols in this study) can be achieved by a widely available NMR method (DOSY) in place of electron microscopy.
- 41Pascal, S. M.; Yamazaki, T.; Singer, A. U.; Kay, L. E.; Forman-Kay, J. D. Structural and dynamic characterization of the phosphotyrosine binding region of an Src homology 2 domain-phosphopeptide complex by NMR relaxation, proton exchange, and chemical shift approaches. Biochemistry 1995, 34, 11353– 11362, DOI: 10.1021/bi00036a008Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXns1Cis7o%253D&md5=775149c2714b25d7fdbb234293b174d2Structural and Dynamic Characterization of the Phosphotyrosine Binding Region of an Src Homology 2 Domain-Phosphopeptide Complex by NMR Relaxation, Proton Exchange, and Chemical Shift ApproachesPascal, Steven M.; Yamazaki, Toshio; Singer, Alex U.; Kay, Lewis E.; Forman-Kay, Julie D.Biochemistry (1995), 34 (36), 11353-62CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Arginine side chains are often involved in protein-protein and protein-nucleic acid interactions. Due to a no. of factors, resonance assignment and detection of NOEs involving the arginine side chains via std. NMR techniques can be difficult. The authors present here an approach to characterization of the interaction between a phosphopeptide (pY1021) and four arginine residues that line the phosphotyrosine-binding pocket of the C-terminal SH2 domain of phospholipase C-γ1 (PLCC SH2). Previously published NOE data provide a partial description of this interaction, including contacts between the aliph. region of Arg 59 and the phosphotyrosine (pTyr) arom. ring. Further characterization has now been accomplished by using 15N and 13C NMR relaxation studies of arginine Νε and Cζ spins, resp., and proton exchange rates of arginine Hε nuclei. Differences between the chem. shifts of the arginine guanidino groups of the free SH2 domain in imidazole and phosphate buffers or in complex with pY1021 have provided insight into specific interactions with the phosphate and the arom. ring of the pTyr. The resulting data are consistent with the most stable hydrogen bonds to phosphate donated by the Arg 39 ε-NH and the two Arg 37 η-NH2 groups and with pTyr arom. ring interactions involving the Arg 39 and possibly the Arg 18 guanidino groups.
- 42Yamazaki, T.; Pascal, S. M.; Singer, A. U.; Forman-Kay, J. D.; Kay, L. E. NMR pulse schemes for the sequence-specific assignment of arginine guanidino 15N and 1H chemical shifts in proteins. J. Am. Chem. Soc. 1995, 117, 3556– 3564, DOI: 10.1021/ja00117a025Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXksVynurg%253D&md5=5678d466099af9086eea72d838cfcb26NMR Pulse Schemes for the Sequence-Specific Assignment of Arginine Guanidino 15N and 1H Chemical Shifts in ProteinsYamazaki, Toshio; Pascal, Steven M.; Singer, Alex U.; Forman-Kay, Julie D.; Kay, Lewis E.Journal of the American Chemical Society (1995), 117 (12), 3556-64CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A family of 2D NMR expts. is presented for the sequence-specific assignment of arginine guanidino 1H and 15N chem. shifts on the basis of transfer of magnetization exclusively by scalar connectivities. Because of the potential for significant exchange with water at the ε and η positions along the side chain of arginine residues, care has been taken to minimize satn. and dephasing of water throughout the course of the pulse schemes. Attempts are made to minimize the effects of chem. exchange due to moderately slow rotation about the Nε-Cζ bond of arginine. The methods are demonstrated on a 1.5 mM sample of the C-terminal SH2 domain from phospholipase-Cγ1 in complex with a 12-residue phosphotyrosyl peptide comprising its high-affinity binding site in the platelet-derived growth factor receptor.
- 43Feng, M.-H.; Philippopoulos, M.; MacKerell, A. D.; Lim, C. Structural Characterization of the Phosphotyrosine Binding Region of a High-Affinity SH2 Domain–Phosphopeptide Complex by Molecular Dynamics Simulation and Chemical Shift Calculations. J. Am. Chem. Soc. 1996, 118, 11265– 11277, DOI: 10.1021/ja961530rGoogle Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmsFOis7g%253D&md5=b0f709df46bea66edad5ba0c738acddeStructural Characterization of the Phosphotyrosine Binding Region of a High-Affinity SH2 Domain-Phosphopeptide Complex by Molecular Dynamics Simulation and Chemical Shift CalculationsFeng, Ming-Hsiang; Philippopoulos, Marios; MacKerell, Alexander D., Jr.; Lim, CarmayJournal of the American Chemical Society (1996), 118 (45), 11265-11277CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Three mol. dynamics simulations of the free, phosphate ion-bound and phosphopeptide-bound C-terminal SH2 domain of phospholipase C-γ1 (PLCC·pY) have been performed to aid in the interpretation of chem. shift data and the elucidation of interat. interactions at the phosphotyrosine (pTyr) binding region. The simulation of the phosphopeptide complex was carried out with newly developed CHARMM force-field parameters for pTyr, optimized against exptl. data and ab initio calcns. The lack of NOEs involving phosphate in the binding pocket had necessitated a chem. shift anal. of the pTyr binding region for a more detailed characterization of the hydrogen bonding interactions involving pTyr. Although most of these interactions are not present in the NMR structure used as the simulation starting point, the system converges early in the simulation to a structure more compatible with the chem. shift data. This is supported by ab initio detn. of the 1Hη and 1Hε chem. shifts of the three arginines (Arg 18, 37, and 39) in the pTyr binding pocket based on the PLCC· pY MD structure, which are in accord with the exptl. values. The simulation structure of the PLCC·pY complex reveals a more complete picture of interat. interactions in the pTyr binding pocket than is possible with current chem. shift and NOE approaches alone, thereby permitting the identification of the primary pTyr-recognition residues. This pattern of interactions is strikingly similar to those of crystal structures of related SH2 domains. The simulations also suggest several alternative interpretations of the chem. shift data to those suggested in the exptl. investigation (Pascal, S. M., et al. Biochem. 1995, 34, 11353). This insight is valuable as the obsd. chem. shifts could result from a no. of possible pictures of interactions. The present study demonstrates that the combination of mol. dynamics simulations and ab initio chem. shift calcns. can enhance the hydrogen-bonding, amino-arom., and aliph.-arom. information content of NOE- and chem.-shift-based protein structures and serve as a complementary tool for the interpretation of chem. shift data at the at. level.
- 44Gargaro, A. R.; Frenkiel, T. A.; Nieto, P. M.; Birdsall, B.; Polshakov, V. I.; Morgan, W. D.; Feeney, J. NMR Detection of arginine-ligand interactions in complexes of lactobacillus casei dihydrofolate reductase. Eur. J. Biochem. 1996, 238, 435– 439, DOI: 10.1111/j.1432-1033.1996.0435z.xGoogle Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjsVKlurY%253D&md5=7326d90843860ab68db86e725710d372NMR detection of arginine-ligand interactions in complexes of Lactobacillus casei dihydrofolate reductaseGargaro, Angelo R.; Frenkiel, Thomas A.; Nieto, Pedro M.; Birdsall, Berry; Polshakov, Vladimir I.; Morgan, William D.; Feeney, JamesEuropean Journal of Biochemistry (1996), 238 (2), 435-439CODEN: EJBCAI; ISSN:0014-2956. (Springer)1H-NMR and 15N-NMR signal assignments have been made for the eight arginine residues in Lactobacillus casei dihydrofolate reductase in its binary complex with methotrexate and in its ternary complex with methotrexate and NADPH. 1H-NMR chem. shifts for the guanidino groups of two of the arginines (Arg57 and Arg43) were sensitive to different modes of binding of the guanidino groups with charged oxygen atoms of the ligands. In the complexes formed with methotrexate, Arg57 showed four nonequivalent NHη proton signals indicating hindered rotation about the Nε-Cζ and Cζ-Nη bonds. The NHη12 and NHη22 protons showed large downfield shifts, which would be expected for a sym. end-on interaction of these protons with the charged oxygen atoms of a carboxylate group in methotrexate. These effects were not obsd. for the complex formed with trimethoprim, which does not contain any carboxylate groups. In the complex formed with NADPH present, Arg43 showed a large downfield chem. shift for its NHε proton and a retardation of its rate of exchange with water. This pattern of deshielding contrasts with that detected for Arg57 and is that expected for a side-on interaction of the guanidino group protons with charged oxygen atoms of the ribose 2'-phosphate group of NADPH.
- 45Nieto, P. M.; Birdsall, B.; Morgan, W. D.; Frenkiel, T. A.; Gargaro, A. R.; Feeney, J. Correlated bond rotations in interactions of arginine residues with ligand carboxylate groups in protein ligand complexes. FEBS Lett. 1997, 405, 16– 20, DOI: 10.1016/s0014-5793(97)00147-6Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXhvFWgs70%253D&md5=8acf19244ec26e694bc65a5dfb5a9cafCorrelated bond rotations in interactions of arginine residues with ligand carboxylate groups in protein ligand complexesNieto, Pedro M.; Birdsall, Berry; Morgan, William D.; Frenkiel, Thomas A.; Gargaro, Angelo R.; Feeney, JamesFEBS Letters (1997), 405 (1), 16-20CODEN: FEBLAL; ISSN:0014-5793. (Elsevier)The 1H/15N HSQC NMR spectra of complexes of Lactobacillus casei dihydrofolate reductase contg. methotrexate recorded at 1°C show four resolved signals for the four NHη protons of the Arg57 residue. This is consistent with hindered rotation in the guanidino group resulting from interactions with the α-carboxylate of methotrexate. Increasing the temp. causes exchange line-broadening and coalescence of signals. Rotation rates for the NεCζ and CζNη bonds have been calcd. from lineshape anal. and from zz-HSQC exchange expts. The interactions between the methotrexate α-carboxylate group and the Arg57 guanidino group decrease the rotation rates for the NεCζ bond by about a factor of 10 and those for the CζNη bonds by more than a factor of 100 with respect to their values in free arginine. Furthermore, the relative rates of rotation about these two bonds are reversed in the protein complexes compared with their values in free arginine indicating that there are concerted rotations about the NεCζ bond of the Arg57 guanidino group and the C'Cα bond of the glutamate α-carboxylate group of methotrexate.
- 46Morgan, W. D.; Birdsall, B.; Nieto, P. M.; Gargaro, A. R.; Feeney, J. 1H/15N HSQC NMR Studies of Ligand Carboxylate Group Interactions with Arginine Residues in Complexes of Brodimoprim Analogues andLactobacilluscaseiDihydrofolate Reductase. Biochemistry 1999, 38, 2127– 2134, DOI: 10.1021/bi982359uGoogle Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXmtFOkug%253D%253D&md5=f2849549a6ce0fd8a3a7473414b00d9c1H/15N HSQC NMR Studies of Ligand Carboxylate Group Interactions with Arginine Residues in Complexes of Brodimoprim Analogs and Lactobacillus casei Dihydrofolate ReductaseMorgan, William D.; Birdsall, Berry; Nieto, Pedro M.; Gargaro, Angelo R.; Feeney, JamesBiochemistry (1999), 38 (7), 2127-2134CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)1H and 15N NMR studies have been undertaken on complexes of Lactobacillus casei dihydrofolate reductase (DHFR) formed with analogs of the antibacterial drug brodimoprim (2,4-diamino-5-(3',5'-dimethoxy-4'-bromobenzyl)pyrimidine) in order to monitor interactions between carboxylate groups on the ligands and basic residues in the protein. These analogs had been designed by computer modeling with carboxylated alkyl chains introduced at the 3'-O position in order to improve their binding properties by making addnl. interactions with basic groups in the protein. Specific interactions between ligand carboxylate groups and the conserved Arg57 residue have been detected in studies of 1H/15N HSQC spectra of complexes of DHFR with both the 4-carboxylate and the 4,6-dicarboxylate brodimoprim analogs. The spectra from both complexes showed four resolved signals for the four NHη protons of the guanidino group of Arg57, and this is consistent with hindered rotation in the guanidino group resulting from interactions with the 4-carboxylate group in each analog. In the spectra of each complex, one of the protons from each of the two NH2 groups and both nitrogens are considerably deshielded compared to the shielding values normally obsd. for such nuclei. This pattern of deshielding is that expected for a sym. end-on interaction of the carboxylate oxygens with the NHη12 and NHη22 guanidino protons. The differences in the degree of deshielding between the complexes of the two structurally similar brodimoprim analogs and the methotrexate indicates that the shielding is very sensitive to geometry, most probably to hydrogen bond lengths. The 1H/15N HSQC spectrum of the DHFR complex with the brodimoprim-6-carboxylate analog does not feature any deshielded Arg NHη protons and this argues against a similar interaction with the Arg57 in this case. It has not proved possible to det. whether the 6-carboxylate in this analog is interacting directly with any residue in the protein. 1H/15N HSQC spectra have been fully assigned for the complexes with the three brodimoprim analogs and chem. shift mapping used to explore interactions in the binding site. The 1H signals of the bound ligands for all three brodimoprim analogs have been assigned. Their 1H chem. shifts were found to be fairly similar in the different complexes indicating that the 2,4-diaminopyrimidine and the benzyl ring are binding in essentially the same binding sites and with the same overall conformation in the different complexes. The rotation rate about the NεCζ bond in the brodimoprim-4,6-dicarboxylate complex with DHFR has been detd. from a zz-HSQC exchange expt., and its value is quite similar to that obsd. in the DHFR·methotrexate complex (24±10 s-1 at 8° and 50±10 s-1 at 15°, resp.). The 1H and 15N chem. shift differences of selected amide and guanidino NH groups, measured between the DHFR complexes, provided further evidence about the interactions involving Arg57 with the 4-carboxylate and 4,6-dicarboxylate brodimoprim analogs.
- 47Polshakov, V. I.; Morgan, W. D.; Birdsall, B.; Feeney, J. Validation of a new restraint docking method for solution structure determinations of protein–ligand complexes. J. Biomol. NMR 1999, 14, 115– 122, DOI: 10.1023/a:1008379225053Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXltVyitLY%253D&md5=56958e9b1d57a1c3bb52b6eafb8d40f9Validation of a new restraint docking method for solution structure determinations of protein-ligand complexesPolshakov, Vladimir I.; Morgan, William D.; Birdsall, Berry; Feeney, JamesJournal of Biomolecular NMR (1999), 14 (2), 115-122CODEN: JBNME9; ISSN:0925-2738. (Kluwer Academic Publishers)A new method is proposed for docking ligands into proteins in cases where an NMR-detd. soln. structure of a related complex is available. The method uses a set of exptl. detd. values for protein-ligand, ligand-ligand, and protein-protein restraints for residues in or near to the binding site, combined with a set of protein-protein restraints involving all the other residues which is taken from the list of restraints previously used to generate the ref. structure of a related complex. This approach differs from ordinary docking methods where the calcn. uses fixed at. coordinates from the ref. structure rather than the restraints used to det. the ref. structure. The binding site residues influenced by replacing the ref. ligand by the new ligand were detd. by monitoring differences in 1H chem. shifts. The method has been validated by showing the excellent agreement between structures of L. casei dihydrofolate reductase. trimetrexate calcd. by conventional methods using a full exptl. detd. set of restraints and those using this new restraint docking method based on an L. casei dihydrofolate reductase.methotrexate ref. structure.
- 48Hajduk, P. J.; Olejniczak, E. T.; Fesik, S. W. One-dimensional relaxation- and diffusion-edited NMR methods for screening compounds that bind to macromolecules. J. Am. Chem. Soc. 1997, 119, 12257– 12261, DOI: 10.1021/ja9715962Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXmsVCh&md5=bcd34f47e094b408b1de34eca7b01051One-Dimensional Relaxation- and Diffusion-Edited NMR Methods for Screening Compounds That Bind to MacromoleculesHajduk, Philip J.; Olejniczak, Edward T.; Fesik, Stephen W.Journal of the American Chemical Society (1997), 119 (50), 12257-12261CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Two one-dimensional 1H NMR techniques for efficiently screening libraries of compds. for binding to macromols. are described that exploit the changes in relaxation or diffusion rates of small mols. which occur upon complex formation. The techniques are demonstrated by detecting ligands that bind to the FK506 binding protein and the catalytic domain of stromelysin in the presence of compds. that do not bind to these proteins. These one-dimensional methods directly detect complex formation between a ligand and a macromol. and thus eliminate false positives often obsd. with other techniques. In addn., since these methods rely on the detection of the uncomplexed compd. rather than the bound ligand or macromol., ligands for macromols. of unlimited size can be detected. Furthermore, active compds. can be directly identified from a mixt., significantly reducing the time and material needed for screening large libraries of compds.
- 49Lin, M.; Shapiro, M. J. mixture analysis in combinatorial chemistry. Application of diffusion-resolved NMR spectroscopy. J. Org. Chem. 1996, 61, 7617– 7619, DOI: 10.1021/jo961315tGoogle Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmtFyqu74%253D&md5=33b61b4d3a74575311b9cc4baef328c8Mixture Analysis in Combinatorial Chemistry. Application of Diffusion-Resolved NMR SpectroscopyLin, Mengfen; Shapiro, Michael J.Journal of Organic Chemistry (1996), 61 (21), 7617-7619CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)Mixt. anal. from a simulated combinatorial split and pool synthesis was accomplished using a new pulse sequence combining LED and TOCSY. Diffusion NMR data were used to decode the individual mols. without the need of prior sepn.
- 50Lin, M.; Shapiro, M. J.; Wareing, J. R. Screening mixtures by affinity NMR. J. Org. Chem. 1997, 62, 8930– 8931, DOI: 10.1021/jo971183jGoogle Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXnvVKmsLg%253D&md5=20152b0dd87f0ba70bdd8ee8ae287f73Screening Mixtures by Affinity NMRLin, Mengfen; Shapiro, Michael J.; Wareing, James R.Journal of Organic Chemistry (1997), 62 (25), 8930-8931CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)A potential advantage of the pulsed field gradient diffusion NMR method over the chem. shift method is that binding is detected by the observation of the ligand NMR spectrum and not by changes in the receptor NMR spectrum. This should allow the expt. to be tuned to the binding affinity of the ligand by changing the relative receptor concn. This premise was tested by making a mixt. of four carboxylic acids and using the hydroquinine 9-phenanthryl ether as the receptor. The binding consts. for the ligands, and the pKa's are given. As the concn. of hydroquinine 9-phenanthryl ether is increased, the other carboxylic acids, sequentially begin to appear in their order of binding affinity. This demonstrates that the system can be tuned to a desired sensitivity level and that mixts. of compds. contg. only a few components can be screened to select components which bind at a desired level of affinity.
- 51Lin, M.; Shapiro, M. J.; Wareing, J. R. Diffusion-Edited NMR–Affinity NMR for Direct Observation of Molecular Interactions. J. Am. Chem. Soc. 1997, 119, 5249– 5250, DOI: 10.1021/ja963654+Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXjt1Churc%253D&md5=af2acda8d726760aa199a7d728f25e0eDiffusion-Edited NMR-Affinity NMR for Direct Observation of Molecular InteractionsLin, Mengfen; Shapiro, Michael J.; Wareing, James R.Journal of the American Chemical Society (1997), 119 (22), 5249-5250CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A small mixt. of compds. can be selectively edited to find and identify components involved in mol. interactions using Pulse Field Gradient NMR techniques. The structure of the ligand can be deduced directly, without requiring sepn. of the mixt., using the LED and TOCSY (DECODES) pulse sequence.
- 52Bleicher, K.; Lin, M.; Shapiro, M. J.; Wareing, J. R. Diffusion Edited NMR: Screening Compound Mixtures by Affinity NMR to Detect Binding Ligands to Vancomycin. J. Org. Chem. 1998, 63, 8486– 8490, DOI: 10.1021/jo9817366Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXmslKjsLk%253D&md5=e9097b03f289b3b553a46cb9819b07c7Diffusion Edited NMR: Screening Compound Mixtures by Affinity NMR to Detect Binding Ligands to VancomycinBleicher, Konrad; Lin, Mengfen; Shapiro, Michael J.; Wareing, James R.Journal of Organic Chemistry (1998), 63 (23), 8486-8490CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)Affinity NMR can be used to produce an edited NMR spectrum that identifies ligands that bind to vancomycin from soln. mixts. contg. nonbinding mols. The Diffusion EnCODEd Spectroscopy (DECODES) expt. performed directly on the same sample can be used to det. the structure of the binding ligands without the need for a phys. sepn. step. The all-D amino acid tetrapeptides DDFA and DDFS, known ligands for vancomycin, were identified in the presence of eight nonbinding tetrapeptides. The bound-ligand signals in the two-dimensional DECODES spectrum are readily identified by comparison with the spectral patterns of the vancomycin cross-peaks in the 2D total correlation spectroscopy and correlation spectroscopy spectra. The screening of soln. mixts. of mols. for direct detection of mol. interactions and structural identification of the interacting ligands provides a powerful new tool to complement methods, such as affinity MS, which rely on the phys. sepn. of mixt. components to identify mol. interactions. The soln. mixts. of compds. for screening by affinity NMR could come from any source where the components are in similar relative amts., including synthesis by combinatorial chem. methods.
- 53Anderson, R. C.; Lin, M.; Shapiro, M. J. Affinity NMR: Decoding DNA Binding. J. Comb. Chem. 1999, 1, 69– 72, DOI: 10.1021/cc980004oGoogle Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXnsFKgt70%253D&md5=f8ec3e35bbe112639a0ec518a7be9a75Affinity NMR: Decoding DNA BindingAnderson, Robert C.; Lin, Mengfen; Shapiro, Michael J.Journal of Combinatorial Chemistry (1999), 1 (1), 69-72CODEN: JCCHFF; ISSN:1520-4766. (American Chemical Society)We have shown that affinity NMR can be used to edit a NMR spectrum so that ligands that have affinity to DNA can be obsd. in the presence of other nonbinding mols. Diffusion encoded spectroscopy (DECODES) can be used to identify the binding ligands. We were able to identify Hoechst 33342 as binding to the Drew-Dickerson dodecamer d(CGCGAATTCGCG)2 in the presence of the nonbinding mols. adenine, adenosine, and thiamine. Affinity NMR appears to be readily applicable to DNA systems for the following reasons. (1) The relaxation rate of the DNA oligonucleotides is favorable, thus the signal intensity loss due to relaxation is not severe. (2) A comparison of the patterns of the DNA cross-peaks upon binding in the two-dimensional total correlation spectroscopy and correlation spectroscopy spectrum are easily performed, and the ligand signals in the two-dimensional DECODES spectrum can be readily identified. (3) The arom. part of the DNA spectrum is devoid of 2D cross-peaks in these correlation spectra, greatly facilitating the interpretation of the bound ligand in the DECODES spectrum.
- 54Mayer, M.; Meyer, B. Characterization of ligand binding by saturation transfer difference NMR spectroscopy. Angew. Chem., Int. Ed. 1999, 38, 1784– 1788, DOI: 10.1002/(sici)1521-3773(19990614)38:12<1784::aid-anie1784>3.0.co;2-qGoogle Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXktVCntb0%253D&md5=03f825af40d22911062112dc6074cf46Characterization of ligand binding by saturation transfer difference NMR spectroscopyMayer, Moriz; Meyer, BerndAngewandte Chemie, International Edition (1999), 38 (12), 1784-1788CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH)The difference between a satn. transfer spectrum and a normal NMR spectrum provides a new and fast method (satn. transfer difference (STD) NMR spectroscopy) to screen compd. libraries for binding activity to proteins. STD NMR spectroscopy of mixts. of potential ligands with as little as 1 nmol of protein yields 1D and 2D NMR spectra that exclusively show signals from mols. with binding affinity. In addn., the ligand's binding epitope is easily identified because ligand residues in direct contact to the protein show much stronger STD signals. For example, the binding specificity of Lewisb-hexasaccharide to Aleulria aurantia agglutinin (AAA) can be mapped to the two fucosyl residues. The efficiency of the STD NMR technique is shown by the binding of N-acetylglucosamine (GlcNAc) to wheat germ agglutinin (WGA).
Cited By
This article is cited by 2 publications.
- Giovanni Montà-González, Félix Sancenón, Ramón Martínez-Máñez, Vicente Martí-Centelles. Purely Covalent Molecular Cages and Containers for Guest Encapsulation. Chemical Reviews 2022, 122
(16)
, 13636-13708. https://doi.org/10.1021/acs.chemrev.2c00198
- Daohui Zhao, Yuqing Wang, Qianwen Su, Libo Li, Jian Zhou. Lysozyme Adsorption on Porous Organic Cages: A Molecular Simulation Study. Langmuir 2020, 36
(41)
, 12299-12308. https://doi.org/10.1021/acs.langmuir.0c02233
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. Illustration of the selected guests (L/S red and D/R blue).
Figure 2
Figure 2. Illustration of the three selected signals of CC3-R.
Figure 3
Figure 3. 1H NMR data for H2 signal of the host with both enantiomers of ligand 1 at several concentrations.
Figure 4
Figure 4. 1H NMR region for the corresponding methyl group signal of ligand 1 at 3 mM and the host at 2.5 mM; (a) rac-1; (b) S-1; (c) R-1; and (d) isolated S-1.
Figure 5
Figure 5. DFT structures for the interaction CC3-R–guest; l-1 and d-1; l-2 and d-2.
References
This article references 54 other publications.
- 1Pu, L. Enantioselective fluorescent sensors: a tale of BINOL. Acc. Chem. Res. 2012, 45, 150– 163, DOI: 10.1021/ar200048d1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtVWhu7%252FJ&md5=47d0d8ac638b37ea88aaaa8b173a7db8Enantioselective Fluorescent Sensors: A Tale of BINOLPu, LinAccounts of Chemical Research (2012), 45 (2), 150-163CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review. The development of automated, high-throughput org. synthesis and screening techniques has created an urgent demand for methods that rapidly det. the enantiomeric compn. of chiral compds. Enantioselective fluorescent sensors offer the potential for real-time, high-sensitivity techniques for detg. enantiomeric data in high-throughput chiral assays. In this Account, the authors describe a range of fluorescent sensors derived from 1,1'-bi-2-naphthol (BINOL), a readily available biaryl compd. with axial chirality. BINOL can be used to construct structurally diverse, chiral fluorescent sensors to carry out highly enantioselective, sensitive recognition of chiral amino alcs., α-hydroxycarboxylic acids, and amino acid derivs. For example, the authors prepd. an (S)-BINOL deriv. whose 3,3'-positions are attached to two chiral amino alc. units, each having two Ph substituents. This compd. shows a fluorescence enhancement of 950-fold in the presence of (S)-mandelic acid but very little change in the presence of (R)-mandelic acid. It also allows the enantiomers of this α-hydroxycarboxylic acid to be visually discriminated by an enantioselective pptn. process. A structurally similar (S)-BINOL-amino alc. mol., but with three rather than two Ph substituents in each of the two amino alc. units, was found to exhibit generally enantioselective fluorescence responses toward structurally diverse α-hydroxycarboxylic acids. The authors further prepd. a pseudoenantiomeric analog of this compd. from (R)-H8BINOL, which has the opposite chiral configuration at both the biaryl center as well as the pendant amino alcs. These two compds. have opposite enantioselectivity in the recognition of a chiral substrate, with distinctly different fluorescence emission wavelengths. By mixing them together, the authors developed a pseudoenantiomeric sensor pair to facilitate chiral assays. Using this pseudoenantiomeric sensor pair allows both the concn. and the enantiomeric compn. of a substrate to be detd. in a single fluorescence measurement. The authors synthesized another compd. by ligating a terpyridine unit to BINOL and found that coordination of a Cu(II) ion to the terpyridine unit almost completely quenched its fluorescence. Displacement of the Cu2+ ion from this complex by chiral amino alcs. leads to enantioselective fluorescence enhancement. This BINOL-terpyridine-Cu(II) complex also exhibits enantioselective gel collapsing in the presence of chiral amino alcs., providing a new visual chiral discrimination method. When light-absorbing conjugated units are attached to the BINOL structure, the resulting multiarmed dendritic mols. show greatly amplified fluorescence responses. Thus, the light harvesting effect of dendrimers can be used to greatly increase the sensitivity of the fluorescent sensors. The progress described here demonstrates that highly enantioselective and sensitive fluorescent sensors can be obtained through a systematic study of the structure-property relation between the sensors and the substrates. These sensors show great potential for the development of rapid assays of chiral org. compds.
- 2Zhang, X.; Yin, J.; Yoon, J. Recent advances in development of chiral fluorescent and colorimetric sensors. Chem. Rev. 2014, 114, 4918– 4959, DOI: 10.1021/cr400568b2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXitVars70%253D&md5=00aec7910eb42651586cabcc683638e8Recent Advances in Development of Chiral Fluorescent and Colorimetric SensorsZhang, Xin; Yin, Jun; Yoon, JuyoungChemical Reviews (Washington, DC, United States) (2014), 114 (9), 4918-4959CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. In this review,the results of these efforts are organized based on the constitutional components and structural features of the sensors including small org. mol. based chiral chemosensors, metal complex based chiral probes, polymer based chiral probes, and nanomaterial based chiral sensors. Attention is also given to the chem. mechanisms involved in the chiral recognition and binding interactions involved in each sensing process.
- 3Simancas, R.; Dari, D.; Velamazan, N.; Navarro, M. T.; Cantin, A.; Jorda, J. L.; Sastre, G.; Corma, A.; Rey, F. Modular organic structure-directing agents for the synthesis of zeolites. Science 2010, 330, 1219– 1222, DOI: 10.1126/science.11962403https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVGrsbfP&md5=2b7a5738846b8dd1930fa1aa133f102fModular Organic Structure-Directing Agents for the Synthesis of ZeolitesSimancas, Raquel; Dari, Djamal; Velamazan, Noemi; Navarro, Maria T.; Cantin, Angel; Jorda, Jose L.; Sastre, German; Corma, Avelino; Rey, FernandoScience (Washington, DC, United States) (2010), 330 (6008), 1219-1222CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Org. structure-directing agents (OSDAs) were used to guide the formation of particular types of pores and channels during the synthesis of zeolites. The authors report that the use of highly versatile OSDAs based on phosphazenes has been successfully introduced for the synthesis of zeolites. This approach has made possible the synthesis of the elusive boggsite zeolite, which is formed by 10- and 12-ring intersecting channels. This topol. and these pore dimensions present interesting opportunities for catalysis in reactions of industrial relevance.
- 4Moliner, M.; Martínez, C.; Corma, A. Synthesis strategies for preparing useful small pore zeolites and zeotypes for gas separations and catalysis. Chem. Mater. 2014, 26, 246– 258, DOI: 10.1021/cm40150954https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXpsVOnuro%253D&md5=18314dab0b5f71260d7ebb504861bbadSynthesis Strategies for Preparing Useful Small Pore Zeolites and Zeotypes for Gas Separations and CatalysisMoliner, Manuel; Martinez, Cristina; Corma, AvelinoChemistry of Materials (2014), 26 (1), 246-258CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)A review; in the last years, the prepn. of small pore zeolites, esp. those presenting large cavities in their structure, have received much attention because they have shown excellent application in catalysis (such as methanol-to-olefins or selective catalytic redn. of NOx) and gas sepns. In the present review, we will focus on the diverse synthetic routes followed to direct the crystn. of small pore zeolites and, on the other hand, the most outstanding applications of those small pore microporous materials.
- 5Moliner, M.; Martínez, C.; Corma, A. Multipore zeolites: synthesis and catalytic applications. Angew. Chem., Int. Ed. 2015, 54, 3560– 3579, DOI: 10.1002/anie.2014063445https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXktVKhu7c%253D&md5=07e31e0d3ff882898b85cdf5918fd327Multipore Zeolites: Synthesis and Catalytic ApplicationsMoliner, Manuel; Martinez, Cristina; Corma, AvelinoAngewandte Chemie, International Edition (2015), 54 (12), 3560-3579CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)A review; in the last few years, important efforts have been made to synthesize so-called "multipore" zeolites, which contain channels of different dimensions within the same cryst. structure. This is a very attractive subject, since the presence of pores of different sizes would favor the preferential diffusion of reactants and products through those different channel systems, allowing unique catalytic activities for specific chem. processes. In this Review we describe the most attractive achievements in the rational synthesis of multipore zeolites, contg. small to extra-large pores, and the improvements reported for relevant chem. processes when these multipore zeolites have been used as catalysts.
- 6Corma, A.; García, H.; Llabrés i Xamena, F. X. Engineering metal organic frameworks for heterogeneous catalysis. Chem. Rev. 2010, 110, 4606– 4655, DOI: 10.1021/cr90039246https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXktFers7Y%253D&md5=ce714f3d4475f4ac2a3d81b13931ccc3Engineering Metal Organic Frameworks for Heterogeneous CatalysisCorma, A.; Garcia, H.; Llabres i Xamena, F. X.Chemical Reviews (Washington, DC, United States) (2010), 110 (8), 4606-4655CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review, including the design of MOFs for catalysis and evaluation of its potential as catalyst as well as catalysis by MOFs with active metal sites, with reactive functional groups, and MOFs as host matrixes or nanometric reaction cavities.
- 7Dhakshinamoorthy, A.; Opanasenko, M.; Čejka, J.; Garcia, H. Metal organic frameworks as heterogeneous catalysts for the production of fine chemicals. Catal. Sci. Technol. 2013, 3, 2509– 2540, DOI: 10.1039/c3cy00350g7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsVSltrzL&md5=a0032e100abaa9a46a7ae87e3c7e2005Metal organic frameworks as heterogeneous catalysts for the production of fine chemicalsDhakshinamoorthy, Amarajothi; Opanasenko, Maksym; Cejka, Jiri; Garcia, HermenegildoCatalysis Science & Technology (2013), 3 (10), 2509-2540CODEN: CSTAGD; ISSN:2044-4753. (Royal Society of Chemistry)A review; this review focuses on the use of metal org. frameworks (MOFs) as catalysts for the synthesis of fine chems. While petrochem. is characterized by gas phase reactions, in which MOFs cannot compete with robust zeolites, MOFs are better suited for liq. phase reactions performed at moderate temps. These are the conditions typically employed for the prodn. of fine chems. characterized by being more complex and diverse mols. of low volatility, but with high added value. For the prepn. of this type of compd., MOFs offer the advantage of wide open porosity in the nanometer scale and a large void vol. In the present review we have summarized the reports that appeared up to early 2013 on the use of MOFs as catalysts in the liq. phase for the prodn. of fine chems., primarily classified according to the type of active site and the functional group formed in the reaction. Prospects for future development in this field are provided in the last section.
- 8Valvekens, P.; Vermoortele, F.; De Vos, D. Metal-organic frameworks as catalysts: the role of metal active sites. Catal. Sci. Technol. 2013, 3, 1435– 1445, DOI: 10.1039/c3cy20813c8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXns1KjsL4%253D&md5=38d269d004b12fe72deddab664c2c606Metal-organic frameworks as catalysts: the role of metal active sitesValvekens, Pieterjan; Vermoortele, Frederik; De Vos, DirkCatalysis Science & Technology (2013), 3 (6), 1435-1445CODEN: CSTAGD; ISSN:2044-4753. (Royal Society of Chemistry)A review. In this perspective first the use of metal-org. frameworks (MOFs) as supports for catalytic functions were critically compared with the possibilities offered by other classes of porous materials. Then the incidental or deliberate formation of active sites in MOF lattices were discuss and some strategies to control the no. and activity of these sites, ultimately resulting in MOF catalysts with improved performance, are reviewed.
- 9Jiang, J.; Yaghi, O. M. Brønsted Acidity in Metal-Organic Frameworks. Chem. Rev. 2015, 115, 6966– 6997, DOI: 10.1021/acs.chemrev.5b002219https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtVKisLvP&md5=c07defc470920b029260a98b76f241e0Bronsted Acidity in Metal-Organic FrameworksJiang, Juncong; Yaghi, Omar M.Chemical Reviews (Washington, DC, United States) (2015), 115 (14), 6966-6997CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review; Bronsted acidity in metal-org. frameworks is discussed.
- 10Seoane, B.; Castellanos, S.; Dikhtiarenko, A.; Kapteijn, F.; Gascon, J. Multi-scale crystal engineering of metal organic frameworks. Coord. Chem. Rev. 2016, 307, 147– 187, DOI: 10.1016/j.ccr.2015.06.00810https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFWlsbbJ&md5=cf82259d468159c7347401a263b435b5Multi-scale crystal engineering of metal organic frameworksSeoane, Beatriz; Castellanos, Sonia; Dikhtiarenko, Alla; Kapteijn, Freek; Gascon, JorgeCoordination Chemistry Reviews (2016), 307 (Part_2), 147-187CODEN: CCHRAM; ISSN:0010-8545. (Elsevier B.V.)A review. During the last decade the no. of publications related to the synthesis of new metal-org. frameworks or coordination polymers has grown almost exponentially. Many of them are devoted to the study of the correlation between the mol. components (linkers and metal nodes or clusters) and the final properties of the resulting structure. Withal, the field of metal-org. frameworks has also witnessed important advances in the development of synthetic tools to control the particle size and shape and to obtain secondary porosity by applying knowledge from three disciplines: crystallog., coordination chem. and chem. engineering. These tools allow for crystal engineering beyond the mol. scale extending over the meso and macro scales, so that certain degree of multi-scale design is already possible. In this manner, MOFs' performance were improved in certain applications by choosing the optimal particle morphol. and dimensions that enhance the materials' properties and/or facilitate their implementation on functional devices. This review highlights the latest advances on MOF crystal engineering, with special emphasis on the meso and macro scales. After discussing some general considerations on the fundamentals of MOF crystn., the authors examine different synthetic approaches developed to tune the MOF particle size, shape and textural properties and the impact this multi-scale MOF crystal engineering showed so far in different applications. Finally, the authors' view on possible future research directions is outlined.
- 11Chen, L.; Luque, R.; Li, Y. Controllable design of tunable nanostructures inside metal-organic frameworks. Chem. Soc. Rev. 2017, 46, 4614– 4630, DOI: 10.1039/c6cs00537c11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXotVeqs7w%253D&md5=7a2d2dc717187c9fabbb83253103d6ceControllable design of tunable nanostructures inside metal-organic frameworksChen, Liyu; Luque, Rafael; Li, YingweiChemical Society Reviews (2017), 46 (15), 4614-4630CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)The controllable encapsulation of nanoentities (such as metal nanoparticles, quantum dots, polyoxometalates, org. and metallorg. mols., biomacromols., and metal-org. polyhedra) into metal-org. frameworks (MOFs) to form composite materials has attracted significant research interest in a variety of fields. These composite materials not only exhibit the properties of both the nanoentities and the MOFs but also display unique and synergistic functionalities. Tuning the sizes, compns., and shapes of nanoentities encapsulated in MOFs enables the final composites to exhibit superior performance to those of the sep. constituents for various applications. In this tutorial review article, we summarized the state-of-the-art development of MOFs contg. encapsulated tunable nanoentities, with special emphasis on the prepn. and synergistic properties of these composites.
- 12Zhang, Y.-B.; Su, J.; Furukawa, H.; Yun, Y.; Gándara, F.; Duong, A.; Zou, X.; Yaghi, O. M. Single-crystal structure of a covalent organic framework. J. Am. Chem. Soc. 2013, 135, 16336– 16339, DOI: 10.1021/ja409033p12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhs1Cktr3N&md5=ea22e0f58fcdda105e732e4fede57bc5Single-Crystal Structure of a Covalent Organic FrameworkZhang, Yue-Biao; Su, Jie; Furukawa, Hiroyasu; Yun, Yifeng; Gandara, Felipe; Duong, Adam; Zou, Xiaodong; Yaghi, Omar M.Journal of the American Chemical Society (2013), 135 (44), 16336-16339CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The crystal structure of a new covalent org. framework, termed COF-320, is detd. by single-crystal 3D electron diffraction using the rotation electron diffraction (RED) method for data collection. The COF crystals are prepd. by an imine condensation of tetra-(4-anilyl)-methane and 4,4'-biphenyldialdehyde in 1,4-dioxane at 120 °C to produce a highly porous 9-fold interwoven diamond net. COF-320 exhibits permanent porosity with a Langmuir surface area of 2400 m2/g and a methane total uptake of 15.0 wt. % (176 cm3/cm3) at 25 °C and 80 bar. The successful detn. of the structure of COF-320 directly from single-crystal samples is an important advance in the development of COF chem.
- 13Díaz, U.; Corma, A. Ordered covalent organic frameworks, COFs and PAFs. From preparation to application. Coord. Chem. Rev. 2016, 311, 85– 124, DOI: 10.1016/j.ccr.2015.12.01013https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XjvFyntA%253D%253D&md5=209868f0abc6cdfc21da8c95b81ca7f5Ordered covalent organic frameworks, COFs and PAFs. From preparation to applicationDiaz, Urbano; Corma, AvelinoCoordination Chemistry Reviews (2016), 311 (), 85-124CODEN: CCHRAM; ISSN:0010-8545. (Elsevier B.V.)Covalent org. frameworks, COFs, and their derived sub-groups based on auto-assembly of exclusively arom. units, PAFs, are emerging into the advanced materials field due to their high free porous vol., structural regularity, robustness, hydrothermal stability, and functional variety. They present high gas uptake capacities and presence of stabilized active functions in the framework. This together with charged low-d. structures combined with their organization through π-conjugated system arrays, open the possibilities of COFs and PAFs to be used as effective materials for adsorption, selective sepn. and catalysis, and in nanotechnol. applications. This review will be focused on self-assembly synthesis mechanisms, physico-chem. characteristics, and applications of this class of promising covalent porous org. structures, out looking their possible future approaches and perspectives.
- 14Das, S.; Heasman, P.; Ben, T.; Qiu, S. Porous Organic Materials: Strategic Design and Structure-Function Correlation. Chem. Rev. 2017, 117, 1515– 1563, DOI: 10.1021/acs.chemrev.6b0043914https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitFOhtLnP&md5=fe6aa70e2a25c48ac78974db574ff0f1Porous organic material strategic design and structure-function correlationDas, Saikat; Heasman, Patrick; Ben, Teng; Qiu, ShilunChemical Reviews (Washington, DC, United States) (2017), 117 (3), 1515-1563CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)Porous org. materials have garnered colossal interest with the scientific fraternity due to their excellent gas sorption performances, catalytic abilities, energy storage capacities, and other intriguing applications. This review encompasses the recent significant breakthroughs and the conventional functions and practices in the field of porous org. materials to find useful applications and imparts a comprehensive understanding of the strategic evolution of the design and synthetic approaches of porous org. materials with tunable characteristics. We present an exhaustive anal. of the design strategies with special emphasis on the topologies of cryst. and amorphous porous org. materials. In addn. to elucidating the structure-function correlation and state-of-the-art applications of porous org. materials, we address the challenges and restrictions that prevent us from realizing porous org. materials with tailored structures and properties for useful applications.
- 15Dawson, R.; Cooper, A. I.; Adams, D. J. Nanoporous organic polymer networks. Prog. Polym. Sci. 2012, 37, 530– 563, DOI: 10.1016/j.progpolymsci.2011.09.00215https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtFeks7o%253D&md5=eeb7686ec5d0d748f906596074fdc8e3Nanoporous organic polymer networksDawson, Robert; Cooper, Andrew I.; Adams, Dave J.Progress in Polymer Science (2012), 37 (4), 530-563CODEN: PRPSB8; ISSN:0079-6700. (Elsevier Ltd.)A review. Nanoporous org. polymer networks are a class of materials consisting solely of the lighter elements in the periodic table. These materials have potential uses in areas such as storage, sepn., and catalysis. Here, we review the different classes of nanoporous polymer networks including covalent org. frameworks, hypercrosslinked polymers, conjugated microporous polymers, and polymers of intrinsic microporosity. The growing variety in synthetic routes to these materials allows a range of different polymer networks to be formed, including cryst. and amorphous structures. It is also possible to incorporate many different kinds of functional groups in a modular fashion. So far, most networks have been examd. from the perspective of gas sorption, and this area is discussed critically and in depth in this review. The use of nanoporous org. polymers for applications such as catalysis and sepns. is an important developing area, and we discuss recent developments as well as highlighting potential future opportunities.
- 16Tan, L.; Tan, B. Hypercrosslinked porous polymer materials: design, synthesis, and applications. Chem. Soc. Rev. 2017, 46, 3322– 3356, DOI: 10.1039/c6cs00851h16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXjtFelu7s%253D&md5=8a267d5ef3aa25a59cce8cfc98e19182Hypercrosslinked porous polymer materials: design, synthesis, and applicationsTan, Liangxiao; Tan, BienChemical Society Reviews (2017), 46 (11), 3322-3356CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. Hypercrosslinked polymers (HCPs) are a series of permanent microporous polymer materials initially reported by Davankov, and have received an increasing level of research interest. In recent years, HCPs have experienced rapid growth due to their remarkable advantages such as diverse synthetic methods, easy functionalization, high surface area, low cost reagents and mild operating conditions. Judicious selection of monomers, appropriate length crosslinkers and optimized reaction conditions yielded a well-developed polymer framework with an adjusted porous topol. Post fabrication of the as developed network facilitates the incorporation of various chem. functionalities that may lead to interesting properties and enhance the selection toward a specific application. To date, numerous HCPs have been prepd. by post-crosslinking polystyrene-based precursors, one-step self-polycondensation or external crosslinking strategies. The advent of these methodologies has prompted researchers to construct well-defined porous polymer networks with customized micromorphol. and functionalities. In this review, we describe not only the basic synthetic principles and strategies of HCPs, but also the advancements in the structural and morphol. study as well as the frontiers of potential applications in energy and environmental fields such as gas storage, carbon capture, removal of pollutants, mol. sepn., catalysis, drug delivery, sensing, etc.
- 17Bojdys, M. J.; Hasell, T.; Severin, N.; Jelfs, K. E.; Rabe, J. P.; Cooper, A. I. Porous organic cage crystals: characterising the porous crystal surface. Chem. Commun. 2012, 48, 11948– 11950, DOI: 10.1039/c2cc36602a17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhslCgtb%252FP&md5=986f5e1ef87ea6017720282e6a2fae1cPorous organic cage crystals: characterizing the porous crystal surfaceBojdys, Michael J.; Hasell, Tom; Severin, Nikolai; Jelfs, Kim E.; Rabe, Juergen P.; Cooper, Andrew I.Chemical Communications (Cambridge, United Kingdom) (2012), 48 (98), 11948-11950CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The characterization of porous cryst. solids often relies on single crystal x-ray diffraction, which does not give direct information about the surface of the material. Here, crystals of a porous org. cage, CC3-R, were studied by at. force microscopy and shown to possess two distinct gas-solid interfaces, proving that the bulk crystal structure is preserved at the porous crystal surface.
- 18Tozawa, T.; Jones, J. T. A.; Swamy, S. I.; Jiang, S.; Adams, D. J.; Shakespeare, S.; Clowes, R.; Bradshaw, D.; Hasell, T.; Chong, S. Y.; Tang, C.; Thompson, S.; Parker, J.; Trewin, A.; Bacsa, J.; Slawin, A. M. Z.; Steiner, A.; Cooper, A. I. Porous organic cages. Nat. Mater. 2009, 8, 973– 978, DOI: 10.1038/nmat254518https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhsVGlt7vP&md5=34d8de0ea8a080d32ce9d3f8b6abc9c3Porous organic cagesTozawa, Tomokazu; Jones, James T. A.; Swamy, Shashikala I.; Jiang, Shan; Adams, Dave J.; Shakespeare, Stephen; Clowes, Rob; Bradshaw, Darren; Hasell, Tom; Chong, Samantha Y.; Tang, Chiu; Thompson, Stephen; Parker, Julia; Trewin, Abbie; Bacsa, John; Slawin, Alexandra M. Z.; Steiner, Alexander; Cooper, Andrew I.Nature Materials (2009), 8 (12), 973-978CODEN: NMAACR; ISSN:1476-1122. (Nature Publishing Group)Porous materials are important in a wide range of applications including mol. sepns. and catalysis. Covalently bonded org. cages can assemble into cryst. microporous materials. The porosity is prefabricated and intrinsic to the mol. cage structure, as opposed to being formed by noncovalent self-assembly of nonporous sub-units. The three-dimensional connectivity between the cage windows is controlled by varying the chem. functionality such that either nonporous or permanently porous assemblies can be produced. Surface areas and gas uptakes for the latter exceed comparable mol. solids. One of the cages can be converted by recrystn. to produce either porous or nonporous polymorphs with apparent Brunauer-Emmett-Teller surface areas of 550 and 23 m2 g-1, resp. These results suggest design principles for responsive porous org. solids and for the modular construction of extended materials from prefabricated mol. pores.
- 19Chen, L.; Reiss, P. S.; Chong, S. Y.; Holden, D.; Jelfs, K. E.; Hasell, T.; Little, M. A.; Kewley, A.; Briggs, M. E.; Stephenson, A.; Thomas, K. M.; Armstrong, J. A.; Bell, J.; Busto, J.; Noel, R.; Liu, J.; Strachan, D. M.; Thallapally, P. K.; Cooper, A. I. Separation of rare gases and chiral molecules by selective binding in porous organic cages. Nat. Mater. 2014, 13, 954– 960, DOI: 10.1038/nmat403519https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFGqu77K&md5=24786fc90209b72d87021b8626a2422fSeparation of rare gases and chiral molecules by selective binding in porous organic cagesChen, Linjiang; Reiss, Paul S.; Chong, Samantha Y.; Holden, Daniel; Jelfs, Kim E.; Hasell, Tom; Little, Marc A.; Kewley, Adam; Briggs, Michael E.; Stephenson, Andrew; Thomas, K. Mark; Armstrong, Jayne A.; Bell, Jon; Busto, Jose; Noel, Raymond; Liu, Jian; Strachan, Denis M.; Thallapally, Praveen K.; Cooper, Andrew I.Nature Materials (2014), 13 (10), 954-960CODEN: NMAACR; ISSN:1476-1122. (Nature Publishing Group)A porous org. cage mol. having unprecedented performance in the solid state for the sepn. of rare gases, such as krypton and xenon, is reported. The selectivity arises from a precise size match between the rare gas and the org. cage cavity, as predicted by mol. simulations. Breakthrough expts. demonstrate real practical potential for the sepn. of krypton, xenon and radon from air at concns. of only a few ppm. Selective binding of chiral org. mols. such as 1-phenylethanol is also demonstrated, suggesting applications in enantioselective sepn.
- 20Briggs, M. E.; Slater, A. G.; Lunt, N.; Jiang, S.; Little, M. A.; Greenaway, R. L.; Hasell, T.; Battilocchio, C.; Ley, S. V.; Cooper, A. I. Dynamic flow synthesis of porous organic cages. Chem. Commun. 2015, 51, 17390– 17393, DOI: 10.1039/c5cc07447a20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhs1eiurnE&md5=03eb47fc8c1bbe1cfcaf09745cc5cf6aDynamic flow synthesis of porous organic cagesBriggs, Michael E.; Slater, Anna G.; Lunt, Neil; Jiang, Shan; Little, Marc A.; Greenaway, Rebecca L.; Hasell, Tom; Battilocchio, Claudio; Ley, Steven V.; Cooper, Andrew I.Chemical Communications (Cambridge, United Kingdom) (2015), 51 (98), 17390-17393CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The dynamic covalent synthesis of two imine-based porous org. cages was successfully transferred from batch to continuous flow. The same flow reactor was then used to scramble the constituents of these two cages in differing ratios to form cage mixts. Preparative HPLC purifn. of one of these mixts. allowed rapid access to a desymmetrised cage mol.
- 21Jiang, S.; Jones, J. T. A.; Hasell, T.; Blythe, C. E.; Adams, D. J.; Trewin, A.; Cooper, A. I. Porous organic molecular solids by dynamic covalent scrambling. Nat. Commun. 2011, 2, 207, DOI: 10.1038/ncomms120721https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3M3htlSrtQ%253D%253D&md5=45e70ff90a4fbf3282cd9d8810d9f5e3Porous organic molecular solids by dynamic covalent scramblingJiang Shan; Jones James T A; Hasell Tom; Blythe Charlotte E; Adams Dave J; Trewin Abbie; Cooper Andrew INature communications (2011), 2 (), 207 ISSN:.The main strategy for constructing porous solids from discrete organic molecules is crystal engineering, which involves forming regular crystalline arrays. Here, we present a chemical approach for desymmetrizing organic cages by dynamic covalent scrambling reactions. This leads to molecules with a distribution of shapes which cannot pack effectively and, hence, do not crystallize, creating porosity in the amorphous solid. The porous properties can be fine tuned by varying the ratio of reagents in the scrambling reaction, and this allows the preparation of materials with high gas selectivities. The molecular engineering of porous amorphous solids complements crystal engineering strategies and may have advantages in some applications, for example, in the compatibilization of functionalities that do not readily cocrystallize.
- 22Little, M. A.; Chong, S. Y.; Schmidtmann, M.; Hasell, T.; Cooper, A. I. Guest control of structure in porous organic cages. Chem. Commun. 2014, 50, 9465– 9468, DOI: 10.1039/c4cc04158e22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFCgtb3E&md5=a039e2cc30ed71763c6f7d4f21eaa7edGuest control of structure in porous organic cagesLittle, Marc A.; Chong, Samantha Y.; Schmidtmann, Marc; Hasell, Tom; Cooper, Andrew I.Chemical Communications (Cambridge, United Kingdom) (2014), 50 (67), 9465-9468CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)Two porous org. cages with different thermodn. polymorphs were induced by co-solvents to interchange their crystal packing modes, thus achieving guest-mediated control over solid-state porosity. In situ crystallog. allows the effect of the co-solvent guests on these structural interconversions to be understood.
- 23Kewley, A.; Stephenson, A.; Chen, L.; Briggs, M. E.; Hasell, T.; Cooper, A. I. Porous organic cages for gas chromatography separations. Chem. Mater. 2015, 27, 3207– 3210, DOI: 10.1021/acs.chemmater.5b0111223https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXlvV2gsLc%253D&md5=885577c1c343019e8f7781f17243ff82Porous Organic Cages for Gas Chromatography SeparationsKewley, Adam; Stephenson, Andrew; Chen, Linjiang; Briggs, Michael E.; Hasell, Tom; Cooper, Andrew I.Chemistry of Materials (2015), 27 (9), 3207-3210CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)Chromatog. is an important process for characterizing and purifying org. mols., but some mixts. remain challenging to sep. The authors report the 1st use of a soln.-processable porous org. cage mol., (I, CC3), as a stationary phase for gas chromatog. Capillary columns were coated with CC3 using a simple, static coating method. These columns are versatile and can sep. alkanes, arom. mixts., and chiral mols. Atomistic simulations indicate that the efficient shape sorting of branched hexane isomers by CC3 results from a combination of kinetic diffusion and surface-interaction effects.
- 24Briggs, M. E.; Cooper, A. I. A perspective on the synthesis, purification, and characterization of porous organic cages. Chem. Mater. 2017, 29, 149– 157, DOI: 10.1021/acs.chemmater.6b0290324https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhsVKrtrnI&md5=297fd2f519911113fc6e20d278c97074A Perspective on the Synthesis, Purification, and Characterization of Porous Organic CagesBriggs, Michael E.; Cooper, Andrew I.Chemistry of Materials (2017), 29 (1), 149-157CODEN: CMATEX; ISSN:0897-4756. (American Chemical Society)A review with authors data. Porous org. cages present many opportunities in functional materials chem., but the synthetic challenges for these mol. solids are somewhat different from those faced in the areas of metal-org. frameworks, covalent-org. frameworks, or porous polymer networks. Here, we highlight the practical methods that we have developed for the design, synthesis, and characterization of imine porous org. cages using CC1 and CC3 as examples. The key points are transferable to other cages, and this perspective should serve as a practical guide to researchers who are new to this field.
- 25Xie, S.-M.; Zhang, J.-H.; Fu, N.; Wang, B.-J.; Chen, L.; Yuan, L.-M. A chiral porous organic cage for molecular recognition using gas chromatography. Anal. Chim. Acta 2016, 903, 156– 163, DOI: 10.1016/j.aca.2015.11.03025https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFantrfN&md5=9975aeabca89d90620731fac00a0f3dcA chiral porous organic cage for molecular recognition using gas chromatographyXie, Sheng-Ming; Zhang, Jun-Hui; Fu, Nan; Wang, Bang-Jin; Chen, Ling; Yuan, Li-MingAnalytica Chimica Acta (2016), 903 (), 156-163CODEN: ACACAM; ISSN:0003-2670. (Elsevier B.V.)Mol. org. cages as shape-persistent org. mols. with permanent and accessible cavities have attracted a lot of interest because of their importance as host-guest systems. Herein, the authors report a chiral porous org. cage (POC) CC9 dild. with a polysiloxane OV-1701 to fabricate a CC9-coated capillary column, which was used for the high-resoln. gas chromatog. sepn. of org. compds., including positional isomers and racemates. On the porous org. cage CC9-coated capillary column, a large no. of racemic compds. such as chiral alcs., esters, ethers and epoxides can be resolved without derivatization. By comparing the chiral recognition ability of the CC9-coated column with the com. available β-DEX 120 column and the POC CC3-R coated column recently reported by the authors' group, the CC9-coated column offered good resoln. during the sepn. of some racemates, that were not sepd. using the β-DEX 120 column or POC CC3-R coated column. Therefore, the CC9-coated column can be complementary to the β-DEX 120 column and CC3-R coated column. The CC9-coated column exhibited great potential for application in the sepn. of positional isomers and enantiomers with great selectivity, high resoln. and good reproducibility.
- 26Zhang, J.-H.; Xie, S.-M.; Chen, L.; Wang, B.-J.; He, P.-G.; Yuan, L.-M. Homochiral porous organic cage with high selectivity for the separation of racemates in gas chromatography. Anal. Chem. 2015, 87, 7817– 7824, DOI: 10.1021/acs.analchem.5b0151226https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFWqtLjO&md5=aab5bfe5ea412eac7aeb48d04e3048edHomochiral Porous Organic Cage with High Selectivity for the Separation of Racemates in Gas ChromatographyZhang, Jun-Hui; Xie, Sheng-Ming; Chen, Ling; Wang, Bang-Jin; He, Pin-Gang; Yuan, Li-MingAnalytical Chemistry (Washington, DC, United States) (2015), 87 (15), 7817-7824CODEN: ANCHAM; ISSN:0003-2700. (American Chemical Society)Porous org. mol. cages as a new type of porous materials have attracted a tremendous attention for their potential applications in recent years. Here the authors report the use of a homochiral porous org. cage (POC) (CC3-R) dild. with a polysiloxane (OV-1701) as a stationary phase for high-resoln. gas chromatog. (GC) with excellent enantioselectivity. A large no. of optical isomers were resolved without derivatization, including chiral alcs., diols, amines, alc. amines, esters, ketones, ethers, halohydrocarbons, org. acids, amino acid Me esters, and sulfoxides. Compared with com. β-DEX 120 and Chirasil-L-Val columns, the CC3-R coated capillary column offered more preeminent enantioselectivity. CC3-R also exhibits good selectivity for the sepn. of isomers, linear alkanes, alcs., and arom. hydrocarbons. The excellent resoln. ability, repeatability, and thermal stability make CC3-R a promising candidate as a novel stationary phase for GC. The study described herein first proves useful com. Also porous org. mol. materials will become more attractive in sepn. science.
- 27Stilbs, P. Fourier transform pulsed-gradient spin-echo studies of molecular diffusion. Prog. Nucl. Magn. Reson. Spectrosc. 1987, 19, 1– 45, DOI: 10.1016/0079-6565(87)80007-927https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXls1Shsg%253D%253D&md5=ef17c6f8b4d2f92752cae2ee3bbf82c1Fourier transform pulsed-gradient spin-echo studies of molecular diffusionStilbs, PeterProgress in Nuclear Magnetic Resonance Spectroscopy (1987), 19 (1), 1-45CODEN: PNMRAT; ISSN:0079-6565.A review with 193 refs. is given on NMR methods (esp. Fourier-transform, pulsed-field-gradient, spin-echo NMR) in measuring mol. diffusion (self-diffusion, mutual diffusion, and multicomponent diffusion) in various types of systems (including liqs., liq. mixts., polymer and polyelectrolyte solns., gels, surfactant aq. solns., and emulsions).
- 28Tyrrell, H. J. V.; Harris, R. K. Diffusion in Liquids; Butterworths: London, 1984.There is no corresponding record for this reference.
- 29Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215– 241, DOI: 10.1007/s00214-007-0310-x29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 30Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650– 654, DOI: 10.1063/1.43895530https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3cXpvFyitA%253D%253D&md5=eb331bad0222adcbe7cad51527273725Self-consistent molecular orbital methods. XX. A basis set for correlated wave functionsKrishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A.Journal of Chemical Physics (1980), 72 (1), 650-4CODEN: JCPSA6; ISSN:0021-9606.A contracted Gaussian basis set (6-311G**) is developed by optimizing exponents and coeffs. at the Moller-Plesset (MP) second-order level for the ground states of first-row atoms. This has a triple split in the valence s and p shells together with a single set of uncontracted polarization functions on each atom. The basis is tested by computing structures and energies for some simple mols. at various levels of MP theory and comparing with expt.
- 31Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford CT, 2009.There is no corresponding record for this reference.
- 32Stejskal, E. O.; Tanner, J. E. Spin Diffusion Measurements: Spin Echoes in the Presence of a Time-Dependent Field Gradient. J. Chem. Phys. 1965, 42, 288– 292, DOI: 10.1063/1.169569032https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF2MXisFOrtQ%253D%253D&md5=c44b1de92c67b6fc3245225937c4ddd0Spin diffusion measurements: spin echoes in the presence of a time-dependent field gradientStejskal, E. O.; Tanner, J. E.Journal of Chemical Physics (1965), 42 (1), 288-92CODEN: JCPSA6; ISSN:0021-9606.A derivation is given of the effect of a time-dependent magnetic field gradient on the spin-echo expt., particularly in the presence of spin diffusion. There are several reasons for preferring certain kinds of time-dependent magnetic field gradients to the more usual steady gradient. If the gradient is reduced during the radio-frequency pulses, H1 need not be particularly large; if the gradient is small at the time of the echo, the echo will be broad and its amplitude easy to measure. Both of these relaxations of restrictions on the measurement of diffusion coeffs. by the spin-echo technique serve to extend its range of applicability. A pulsed gradient can be recommended when it is crit. to define the precise time period over which diffusion is being measured. The theoretical expression derived was verified exptl. for several choices of time-dependent magnetic field gradient. An app. is described suitable for the production of pulsed gradients with amplitudes as large as 100 gauss cm. The diffusion coeff. of dry glycerol at 26° ± 1° is (2.5 ± 0.2) × 10-3 cm.2/sec., a value smaller than can ordinarily be measured by the steady gradient method.
- 33Tanner, J. E. Use of the stimulated echo in NMR diffusion studies. J. Chem. Phys. 1970, 52, 2523– 2526, DOI: 10.1063/1.167333633https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXptlGgsw%253D%253D&md5=7390bc823359258bfb885733b25de844Use of the stimulated echo in NMR diffusion studiesTanner, John E.Journal of Chemical Physics (1970), 52 (5), 2523-6CODEN: JCPSA6; ISSN:0021-9606.The stimulated echo in a three-radio-frequency-pulse expt. is useful in extending the range of measurement of diffusion coeffs. to more viscous substances or the measurement of barrier sepns. to wider spacings in systems where the diffusing substance has T1 > T2. The spin-echo attenuation due to self-diffusion is derived for the general case of a time-dependent field gradient, and the result is found exptl. to be correct for the special case of a field gradient applied in 2 equal, square pulse.
- 34Johnson, C. S. Diffusion ordered nuclear magnetic resonance spectroscopy: principles and applications. Magn. Reson. Spectrosc. 1999, 34, 203– 256, DOI: 10.1016/s0079-6565(99)00003-534https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXivVOltbc%253D&md5=956048c606a7721394fc41687da69820Diffusion ordered nuclear magnetic resonance spectroscopy: principles and applicationsJohnson, C. S., Jr.Progress in Nuclear Magnetic Resonance Spectroscopy (1999), 34 (3,4), 203-256CODEN: PNMRAT; ISSN:0079-6565. (Elsevier Science B.V.)A review with 146 refs.
- 35Chen, A.; Wu, D.; Johnson, C. S. Determination of molecular weight distributions for polymers by diffusion-ordered NMR. J. Am. Chem. Soc. 1995, 117, 7965– 7970, DOI: 10.1021/ja00135a01535https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmvVegsbc%253D&md5=ed023c8af0c849c7a607da6fc11a93edDetermination of Molecular Weight Distributions for Polymers by Diffusion-Ordered NMRChen, Aidi; Wu, Donghui; Johnson, Charles S., Jr.Journal of the American Chemical Society (1995), 117 (30), 7965-70CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Diffusion-ordered NMR spectroscopy, a technique based on pulsed field gradient NMR (PFGNMR), was used to characterize mol. wt. distributions for samples of poly(ethylene oxide) in D2O. The distribution of diffusion coeffs. G(D) was obtained by anal. of PFGNMR data with a modified version of the well-known constrained regularization program CONTIN. Regularization was considerably improved for broad distributions by setting the wts. cm in CONTIN's quadrature formula equal to (Dm/Dmax)xm, where Dmax is the diffusion coeff. corresponding to the max. in G(Dm)Dm and m is an index. Here xm varies linearly from +2 to -2 with log(Dm) across the distribution. This amts. to enhancing low amplitude regions of G(D)D during anal. The estd. distribution was then converted to the mass weighted distribution of mol. wts. by means of the relation D = 10-7.62 M-0.62 (with D in units of m2 s-1) obtained from expts. on monodisperse ref. stds. In this study spin relaxation rates were independent of mol. wts. and intermol. averaging effects were insignificant. As an illustration, mol. wt. distributions were detd. for two broadly distributed samples. The no. and wt. av. mol. wts. and the polydispersities agreed well with values provided by the manufacturer when the PFGNMR data sets had signal-to-noise ratios greater than 500.
- 36Avram, L.; Cohen, Y. Diffusion NMR of molecular cages and capsules. Chem. Soc. Rev. 2015, 44, 586– 602, DOI: 10.1039/c4cs00197d36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtlCjurzL&md5=8f40f901cfd7bac40cb022bda464359fDiffusion NMR of molecular cages and capsulesAvram, Liat; Cohen, YoramChemical Society Reviews (2015), 44 (2), 586-602CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. In the last decade diffusion NMR and diffusion ordered spectroscopy (DOSY) have become important anal. tools for the characterization of supramol. systems in soln. Diffusion NMR can be used to glean information on the (effective) size and shape of mol. species, as well as to probe inter-mol. interactions and can be used to est. the assocn. const. (Ka) of a complex. In addn., the diffusion coeff., as obtained from diffusion NMR, is a much more intuitive parameter than the chem. shift for probing self-assocn., aggregation and inter-mol. interactions. The diffusion coeff. may be an even more important anal. parameter in systems in which the formed supramol. entity has the same symmetry as its building units, when there is a large change in the mol. wt., where many mol. species are involved in the formation of the supramol. systems, and when proton transfer may occur which, in turn, may affect the chem. shift. Some of the self-assembled mol. capsules and cages prepd. in the last decade represent such supramol. systems and in the present review, following a short introduction on diffusion NMR, we survey the contribution of diffusion NMR and DOSY in the field of mol. containers and capsules. We will 1st focus on the role played by diffusion NMR in the field of H-bond driven self-assembled capsules. We then survey the contributions of diffusion NMR and DOSY to the study and characterization of metal-ligand cages and capsules. Finally, we describe a few recent applications of diffusion NMR in the field of hydrophobic, electrostatic and covalent containers.
- 37Åslund, I.; Nowacka, A.; Nilsson, M.; Topgaard, D. Filter-exchange PGSE NMR determination of cell membrane permeability. J. Magn. Reson. 2009, 200, 291– 295, DOI: 10.1016/j.jmr.2009.07.01537https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BD1MnisVGnuw%253D%253D&md5=cf19b3e4a54adf6709dee41ae77a48dbFilter-exchange PGSE NMR determination of cell membrane permeabilityAslund Ingrid; Nowacka Agnieszka; Nilsson Markus; Topgaard DanielJournal of magnetic resonance (San Diego, Calif. : 1997) (2009), 200 (2), 291-5 ISSN:.A new PGSE NMR sequence is introduced for measuring diffusive transport across the plasma membrane of living cells. A "diffusion filter" and a variable mixing time precedes a standard PGSE block for diffusion encoding of the NMR signal. The filter is a PGSE block optimized for selectively removing the magnetization of the extracellular water. With increasing mixing time the intra- and extracellular components approach their equilibrium fractional populations. The rate of exchange can be measured using only a few minutes of instrument time. Water exchange over the plasma membrane of starved yeast cells is studied in the temperature range +5 to +32 degrees C.
- 38Occhipinti, P.; Griffiths, P. C. Quantifying diffusion in mucosal systems by pulsed-gradient spin-echo NMR. Adv. Drug Delivery Rev. 2008, 60, 1570– 1582, DOI: 10.1016/j.addr.2008.08.00638https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhsVSjsL%252FM&md5=9d3a2f6cec6bfcac0575c7d0c61b1583Quantifying diffusion in mucosal systems by pulsed-gradient spin-echo NMROcchipinti, Paola; Griffiths, Peter C.Advanced Drug Delivery Reviews (2008), 60 (15), 1570-1582CODEN: ADDREP; ISSN:0169-409X. (Elsevier B.V.)A review. Mucus, a thick and slimy secretion produced by submucosal cells, covers many epithelial surfaces in mammalian organs and prevents foreign particles that enter the body from accessing cells. However, the mucus layer also represents a potential barrier to the efficient delivery of nano-sized drug delivery systems (polyplexes, lipoplexes, particles) to the underlying mucosal epithelium. Many studies have considered the ability of nano-sized particles and polymers to diffuse within the mucosal network using a range of different techniques, including multiple-particle tracking (MPT), diffusion chamber studies and fluorescence recovery after photobleaching (FRAP). This review highlights the current understanding of the interaction of the diffusion of nano-sized structures within mucosal networks. Moreover, this article presents an introduction to pulsed-gradient spin-echo NMR (PGSE-NMR), a potential new tool to investigate the mobility of mol. species through mucosal networks and related biol. gels.
- 39Marega, R.; Aroulmoji, V.; Dinon, F.; Vaccari, L.; Giordani, S.; Bianco, A.; Murano, E.; Prato, M. Diffusion-ordered NMR spectroscopy in the structural characterization of functionalized carbon nanotubes. J. Am. Chem. Soc. 2009, 131, 9086– 9093, DOI: 10.1021/ja902728w39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXmtlWhu7Y%253D&md5=2d08688be42573ccff4023551e4b38a3Diffusion-Ordered NMR Spectroscopy in the Structural Characterization of Functionalized Carbon NanotubesMarega, Riccardo; Aroulmoji, Vincent; Dinon, Francesca; Vaccari, Lisa; Giordani, Silvia; Bianco, Alberto; Murano, Erminio; Prato, MaurizioJournal of the American Chemical Society (2009), 131 (25), 9086-9093CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The emerging applications of functionalized C nanotubes (CNTs) in various research domains necessitate the use of many different anal. techniques to confirm their structural modifications in a fast and reliable manner. Thus far, NMR spectroscopy was not among the main tools for characterization of organically modified C nanostructures. 1H anal. is limited because the signals in these derivs. are typically weak and broad, resulting in uncertainties of a few ppm, and because of the strong interference of residual solvent signals. To overcome these limitations, the authors studied the applicability of 1H NMR spectroscopy based on gradient-edited diffusion pulse sequences (1-dimensional diffusion-ordered spectroscopy, DOSY) in the characterization of CNT derivs. In general, diffusion NMR expts. allow the sepn. of NMR signals of different species present in a mixt., according to their own diffusion coeffs., merging spectroscopy information with size anal. A selected set of CNT derivs. was synthesized and analyzed using 1-dimensional DOSY expts. by applying strong magnetic field gradients (up to 42.6 G cm-1). Colorimetric tests (i.e., Kaiser test) and TGA anal. support the NMR findings, which are related to isolated and/or bundled short SWNTs, from TEM and AFM characterization. The overall results show that the diffusion-based NMR spectroscopy is a fast and promising approach for the characterization of covalently modified CNT derivs.
- 40Canzi, G.; Mrse, A. A.; Kubiak, C. P. Diffusion-ordered NMR spectroscopy as a reliable alternative to TEM for determining the size of gold nanoparticles in organic solutions. J. Phys. Chem. C 2011, 115, 7972– 7978, DOI: 10.1021/jp200855740https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXktFelsbk%253D&md5=4c42a9f46d6d381a10a00381fe8c1ac3Diffusion-Ordered NMR Spectroscopy as a Reliable Alternative to TEM for Determining the Size of Gold Nanoparticles in Organic SolutionsCanzi, Gabriele; Mrse, Anthony A.; Kubiak, Clifford P.Journal of Physical Chemistry C (2011), 115 (16), 7972-7978CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Diffusion-ordered spectroscopy is used to det. gold nanoparticle sizes. Traditional characterization of nanoparticles has centered on imaging by electron microscopy and plasmon resonance absorption in UV-visible electronic spectra. The authors present a convenient method to characterize gold nanoparticles using diffusion-ordered NMR spectroscopy (DOSY). 2-dimensional DOSY NMR is used to calc. diffusion consts. and the diam. of solubilized gold nanoparticles capped with 1-dodecanethiol (C12) or 1-octanethiol (C8) in three deuterated solvents. The distributions of nanoparticle sizes strongly correlate with TEM image anal. C12 and C8 capped nanoparticle sizes are 4.6 and 2.7 nm by TEM as compared to ests. of 4.6 ± 0.3 and 2.5 ± 0.1 nm based on 2-dimensional DOSY NMR data. Reliable size characterization of nanoparticles with NMR active nuclei (1H in this study) in their protective groups (alkane thiols in this study) can be achieved by a widely available NMR method (DOSY) in place of electron microscopy.
- 41Pascal, S. M.; Yamazaki, T.; Singer, A. U.; Kay, L. E.; Forman-Kay, J. D. Structural and dynamic characterization of the phosphotyrosine binding region of an Src homology 2 domain-phosphopeptide complex by NMR relaxation, proton exchange, and chemical shift approaches. Biochemistry 1995, 34, 11353– 11362, DOI: 10.1021/bi00036a00841https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXns1Cis7o%253D&md5=775149c2714b25d7fdbb234293b174d2Structural and Dynamic Characterization of the Phosphotyrosine Binding Region of an Src Homology 2 Domain-Phosphopeptide Complex by NMR Relaxation, Proton Exchange, and Chemical Shift ApproachesPascal, Steven M.; Yamazaki, Toshio; Singer, Alex U.; Kay, Lewis E.; Forman-Kay, Julie D.Biochemistry (1995), 34 (36), 11353-62CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Arginine side chains are often involved in protein-protein and protein-nucleic acid interactions. Due to a no. of factors, resonance assignment and detection of NOEs involving the arginine side chains via std. NMR techniques can be difficult. The authors present here an approach to characterization of the interaction between a phosphopeptide (pY1021) and four arginine residues that line the phosphotyrosine-binding pocket of the C-terminal SH2 domain of phospholipase C-γ1 (PLCC SH2). Previously published NOE data provide a partial description of this interaction, including contacts between the aliph. region of Arg 59 and the phosphotyrosine (pTyr) arom. ring. Further characterization has now been accomplished by using 15N and 13C NMR relaxation studies of arginine Νε and Cζ spins, resp., and proton exchange rates of arginine Hε nuclei. Differences between the chem. shifts of the arginine guanidino groups of the free SH2 domain in imidazole and phosphate buffers or in complex with pY1021 have provided insight into specific interactions with the phosphate and the arom. ring of the pTyr. The resulting data are consistent with the most stable hydrogen bonds to phosphate donated by the Arg 39 ε-NH and the two Arg 37 η-NH2 groups and with pTyr arom. ring interactions involving the Arg 39 and possibly the Arg 18 guanidino groups.
- 42Yamazaki, T.; Pascal, S. M.; Singer, A. U.; Forman-Kay, J. D.; Kay, L. E. NMR pulse schemes for the sequence-specific assignment of arginine guanidino 15N and 1H chemical shifts in proteins. J. Am. Chem. Soc. 1995, 117, 3556– 3564, DOI: 10.1021/ja00117a02542https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXksVynurg%253D&md5=5678d466099af9086eea72d838cfcb26NMR Pulse Schemes for the Sequence-Specific Assignment of Arginine Guanidino 15N and 1H Chemical Shifts in ProteinsYamazaki, Toshio; Pascal, Steven M.; Singer, Alex U.; Forman-Kay, Julie D.; Kay, Lewis E.Journal of the American Chemical Society (1995), 117 (12), 3556-64CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A family of 2D NMR expts. is presented for the sequence-specific assignment of arginine guanidino 1H and 15N chem. shifts on the basis of transfer of magnetization exclusively by scalar connectivities. Because of the potential for significant exchange with water at the ε and η positions along the side chain of arginine residues, care has been taken to minimize satn. and dephasing of water throughout the course of the pulse schemes. Attempts are made to minimize the effects of chem. exchange due to moderately slow rotation about the Nε-Cζ bond of arginine. The methods are demonstrated on a 1.5 mM sample of the C-terminal SH2 domain from phospholipase-Cγ1 in complex with a 12-residue phosphotyrosyl peptide comprising its high-affinity binding site in the platelet-derived growth factor receptor.
- 43Feng, M.-H.; Philippopoulos, M.; MacKerell, A. D.; Lim, C. Structural Characterization of the Phosphotyrosine Binding Region of a High-Affinity SH2 Domain–Phosphopeptide Complex by Molecular Dynamics Simulation and Chemical Shift Calculations. J. Am. Chem. Soc. 1996, 118, 11265– 11277, DOI: 10.1021/ja961530r43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmsFOis7g%253D&md5=b0f709df46bea66edad5ba0c738acddeStructural Characterization of the Phosphotyrosine Binding Region of a High-Affinity SH2 Domain-Phosphopeptide Complex by Molecular Dynamics Simulation and Chemical Shift CalculationsFeng, Ming-Hsiang; Philippopoulos, Marios; MacKerell, Alexander D., Jr.; Lim, CarmayJournal of the American Chemical Society (1996), 118 (45), 11265-11277CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Three mol. dynamics simulations of the free, phosphate ion-bound and phosphopeptide-bound C-terminal SH2 domain of phospholipase C-γ1 (PLCC·pY) have been performed to aid in the interpretation of chem. shift data and the elucidation of interat. interactions at the phosphotyrosine (pTyr) binding region. The simulation of the phosphopeptide complex was carried out with newly developed CHARMM force-field parameters for pTyr, optimized against exptl. data and ab initio calcns. The lack of NOEs involving phosphate in the binding pocket had necessitated a chem. shift anal. of the pTyr binding region for a more detailed characterization of the hydrogen bonding interactions involving pTyr. Although most of these interactions are not present in the NMR structure used as the simulation starting point, the system converges early in the simulation to a structure more compatible with the chem. shift data. This is supported by ab initio detn. of the 1Hη and 1Hε chem. shifts of the three arginines (Arg 18, 37, and 39) in the pTyr binding pocket based on the PLCC· pY MD structure, which are in accord with the exptl. values. The simulation structure of the PLCC·pY complex reveals a more complete picture of interat. interactions in the pTyr binding pocket than is possible with current chem. shift and NOE approaches alone, thereby permitting the identification of the primary pTyr-recognition residues. This pattern of interactions is strikingly similar to those of crystal structures of related SH2 domains. The simulations also suggest several alternative interpretations of the chem. shift data to those suggested in the exptl. investigation (Pascal, S. M., et al. Biochem. 1995, 34, 11353). This insight is valuable as the obsd. chem. shifts could result from a no. of possible pictures of interactions. The present study demonstrates that the combination of mol. dynamics simulations and ab initio chem. shift calcns. can enhance the hydrogen-bonding, amino-arom., and aliph.-arom. information content of NOE- and chem.-shift-based protein structures and serve as a complementary tool for the interpretation of chem. shift data at the at. level.
- 44Gargaro, A. R.; Frenkiel, T. A.; Nieto, P. M.; Birdsall, B.; Polshakov, V. I.; Morgan, W. D.; Feeney, J. NMR Detection of arginine-ligand interactions in complexes of lactobacillus casei dihydrofolate reductase. Eur. J. Biochem. 1996, 238, 435– 439, DOI: 10.1111/j.1432-1033.1996.0435z.x44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XjsVKlurY%253D&md5=7326d90843860ab68db86e725710d372NMR detection of arginine-ligand interactions in complexes of Lactobacillus casei dihydrofolate reductaseGargaro, Angelo R.; Frenkiel, Thomas A.; Nieto, Pedro M.; Birdsall, Berry; Polshakov, Vladimir I.; Morgan, William D.; Feeney, JamesEuropean Journal of Biochemistry (1996), 238 (2), 435-439CODEN: EJBCAI; ISSN:0014-2956. (Springer)1H-NMR and 15N-NMR signal assignments have been made for the eight arginine residues in Lactobacillus casei dihydrofolate reductase in its binary complex with methotrexate and in its ternary complex with methotrexate and NADPH. 1H-NMR chem. shifts for the guanidino groups of two of the arginines (Arg57 and Arg43) were sensitive to different modes of binding of the guanidino groups with charged oxygen atoms of the ligands. In the complexes formed with methotrexate, Arg57 showed four nonequivalent NHη proton signals indicating hindered rotation about the Nε-Cζ and Cζ-Nη bonds. The NHη12 and NHη22 protons showed large downfield shifts, which would be expected for a sym. end-on interaction of these protons with the charged oxygen atoms of a carboxylate group in methotrexate. These effects were not obsd. for the complex formed with trimethoprim, which does not contain any carboxylate groups. In the complex formed with NADPH present, Arg43 showed a large downfield chem. shift for its NHε proton and a retardation of its rate of exchange with water. This pattern of deshielding contrasts with that detected for Arg57 and is that expected for a side-on interaction of the guanidino group protons with charged oxygen atoms of the ribose 2'-phosphate group of NADPH.
- 45Nieto, P. M.; Birdsall, B.; Morgan, W. D.; Frenkiel, T. A.; Gargaro, A. R.; Feeney, J. Correlated bond rotations in interactions of arginine residues with ligand carboxylate groups in protein ligand complexes. FEBS Lett. 1997, 405, 16– 20, DOI: 10.1016/s0014-5793(97)00147-645https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXhvFWgs70%253D&md5=8acf19244ec26e694bc65a5dfb5a9cafCorrelated bond rotations in interactions of arginine residues with ligand carboxylate groups in protein ligand complexesNieto, Pedro M.; Birdsall, Berry; Morgan, William D.; Frenkiel, Thomas A.; Gargaro, Angelo R.; Feeney, JamesFEBS Letters (1997), 405 (1), 16-20CODEN: FEBLAL; ISSN:0014-5793. (Elsevier)The 1H/15N HSQC NMR spectra of complexes of Lactobacillus casei dihydrofolate reductase contg. methotrexate recorded at 1°C show four resolved signals for the four NHη protons of the Arg57 residue. This is consistent with hindered rotation in the guanidino group resulting from interactions with the α-carboxylate of methotrexate. Increasing the temp. causes exchange line-broadening and coalescence of signals. Rotation rates for the NεCζ and CζNη bonds have been calcd. from lineshape anal. and from zz-HSQC exchange expts. The interactions between the methotrexate α-carboxylate group and the Arg57 guanidino group decrease the rotation rates for the NεCζ bond by about a factor of 10 and those for the CζNη bonds by more than a factor of 100 with respect to their values in free arginine. Furthermore, the relative rates of rotation about these two bonds are reversed in the protein complexes compared with their values in free arginine indicating that there are concerted rotations about the NεCζ bond of the Arg57 guanidino group and the C'Cα bond of the glutamate α-carboxylate group of methotrexate.
- 46Morgan, W. D.; Birdsall, B.; Nieto, P. M.; Gargaro, A. R.; Feeney, J. 1H/15N HSQC NMR Studies of Ligand Carboxylate Group Interactions with Arginine Residues in Complexes of Brodimoprim Analogues andLactobacilluscaseiDihydrofolate Reductase. Biochemistry 1999, 38, 2127– 2134, DOI: 10.1021/bi982359u46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXmtFOkug%253D%253D&md5=f2849549a6ce0fd8a3a7473414b00d9c1H/15N HSQC NMR Studies of Ligand Carboxylate Group Interactions with Arginine Residues in Complexes of Brodimoprim Analogs and Lactobacillus casei Dihydrofolate ReductaseMorgan, William D.; Birdsall, Berry; Nieto, Pedro M.; Gargaro, Angelo R.; Feeney, JamesBiochemistry (1999), 38 (7), 2127-2134CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)1H and 15N NMR studies have been undertaken on complexes of Lactobacillus casei dihydrofolate reductase (DHFR) formed with analogs of the antibacterial drug brodimoprim (2,4-diamino-5-(3',5'-dimethoxy-4'-bromobenzyl)pyrimidine) in order to monitor interactions between carboxylate groups on the ligands and basic residues in the protein. These analogs had been designed by computer modeling with carboxylated alkyl chains introduced at the 3'-O position in order to improve their binding properties by making addnl. interactions with basic groups in the protein. Specific interactions between ligand carboxylate groups and the conserved Arg57 residue have been detected in studies of 1H/15N HSQC spectra of complexes of DHFR with both the 4-carboxylate and the 4,6-dicarboxylate brodimoprim analogs. The spectra from both complexes showed four resolved signals for the four NHη protons of the guanidino group of Arg57, and this is consistent with hindered rotation in the guanidino group resulting from interactions with the 4-carboxylate group in each analog. In the spectra of each complex, one of the protons from each of the two NH2 groups and both nitrogens are considerably deshielded compared to the shielding values normally obsd. for such nuclei. This pattern of deshielding is that expected for a sym. end-on interaction of the carboxylate oxygens with the NHη12 and NHη22 guanidino protons. The differences in the degree of deshielding between the complexes of the two structurally similar brodimoprim analogs and the methotrexate indicates that the shielding is very sensitive to geometry, most probably to hydrogen bond lengths. The 1H/15N HSQC spectrum of the DHFR complex with the brodimoprim-6-carboxylate analog does not feature any deshielded Arg NHη protons and this argues against a similar interaction with the Arg57 in this case. It has not proved possible to det. whether the 6-carboxylate in this analog is interacting directly with any residue in the protein. 1H/15N HSQC spectra have been fully assigned for the complexes with the three brodimoprim analogs and chem. shift mapping used to explore interactions in the binding site. The 1H signals of the bound ligands for all three brodimoprim analogs have been assigned. Their 1H chem. shifts were found to be fairly similar in the different complexes indicating that the 2,4-diaminopyrimidine and the benzyl ring are binding in essentially the same binding sites and with the same overall conformation in the different complexes. The rotation rate about the NεCζ bond in the brodimoprim-4,6-dicarboxylate complex with DHFR has been detd. from a zz-HSQC exchange expt., and its value is quite similar to that obsd. in the DHFR·methotrexate complex (24±10 s-1 at 8° and 50±10 s-1 at 15°, resp.). The 1H and 15N chem. shift differences of selected amide and guanidino NH groups, measured between the DHFR complexes, provided further evidence about the interactions involving Arg57 with the 4-carboxylate and 4,6-dicarboxylate brodimoprim analogs.
- 47Polshakov, V. I.; Morgan, W. D.; Birdsall, B.; Feeney, J. Validation of a new restraint docking method for solution structure determinations of protein–ligand complexes. J. Biomol. NMR 1999, 14, 115– 122, DOI: 10.1023/a:100837922505347https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXltVyitLY%253D&md5=56958e9b1d57a1c3bb52b6eafb8d40f9Validation of a new restraint docking method for solution structure determinations of protein-ligand complexesPolshakov, Vladimir I.; Morgan, William D.; Birdsall, Berry; Feeney, JamesJournal of Biomolecular NMR (1999), 14 (2), 115-122CODEN: JBNME9; ISSN:0925-2738. (Kluwer Academic Publishers)A new method is proposed for docking ligands into proteins in cases where an NMR-detd. soln. structure of a related complex is available. The method uses a set of exptl. detd. values for protein-ligand, ligand-ligand, and protein-protein restraints for residues in or near to the binding site, combined with a set of protein-protein restraints involving all the other residues which is taken from the list of restraints previously used to generate the ref. structure of a related complex. This approach differs from ordinary docking methods where the calcn. uses fixed at. coordinates from the ref. structure rather than the restraints used to det. the ref. structure. The binding site residues influenced by replacing the ref. ligand by the new ligand were detd. by monitoring differences in 1H chem. shifts. The method has been validated by showing the excellent agreement between structures of L. casei dihydrofolate reductase. trimetrexate calcd. by conventional methods using a full exptl. detd. set of restraints and those using this new restraint docking method based on an L. casei dihydrofolate reductase.methotrexate ref. structure.
- 48Hajduk, P. J.; Olejniczak, E. T.; Fesik, S. W. One-dimensional relaxation- and diffusion-edited NMR methods for screening compounds that bind to macromolecules. J. Am. Chem. Soc. 1997, 119, 12257– 12261, DOI: 10.1021/ja971596248https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXmsVCh&md5=bcd34f47e094b408b1de34eca7b01051One-Dimensional Relaxation- and Diffusion-Edited NMR Methods for Screening Compounds That Bind to MacromoleculesHajduk, Philip J.; Olejniczak, Edward T.; Fesik, Stephen W.Journal of the American Chemical Society (1997), 119 (50), 12257-12261CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Two one-dimensional 1H NMR techniques for efficiently screening libraries of compds. for binding to macromols. are described that exploit the changes in relaxation or diffusion rates of small mols. which occur upon complex formation. The techniques are demonstrated by detecting ligands that bind to the FK506 binding protein and the catalytic domain of stromelysin in the presence of compds. that do not bind to these proteins. These one-dimensional methods directly detect complex formation between a ligand and a macromol. and thus eliminate false positives often obsd. with other techniques. In addn., since these methods rely on the detection of the uncomplexed compd. rather than the bound ligand or macromol., ligands for macromols. of unlimited size can be detected. Furthermore, active compds. can be directly identified from a mixt., significantly reducing the time and material needed for screening large libraries of compds.
- 49Lin, M.; Shapiro, M. J. mixture analysis in combinatorial chemistry. Application of diffusion-resolved NMR spectroscopy. J. Org. Chem. 1996, 61, 7617– 7619, DOI: 10.1021/jo961315t49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmtFyqu74%253D&md5=33b61b4d3a74575311b9cc4baef328c8Mixture Analysis in Combinatorial Chemistry. Application of Diffusion-Resolved NMR SpectroscopyLin, Mengfen; Shapiro, Michael J.Journal of Organic Chemistry (1996), 61 (21), 7617-7619CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)Mixt. anal. from a simulated combinatorial split and pool synthesis was accomplished using a new pulse sequence combining LED and TOCSY. Diffusion NMR data were used to decode the individual mols. without the need of prior sepn.
- 50Lin, M.; Shapiro, M. J.; Wareing, J. R. Screening mixtures by affinity NMR. J. Org. Chem. 1997, 62, 8930– 8931, DOI: 10.1021/jo971183j50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXnvVKmsLg%253D&md5=20152b0dd87f0ba70bdd8ee8ae287f73Screening Mixtures by Affinity NMRLin, Mengfen; Shapiro, Michael J.; Wareing, James R.Journal of Organic Chemistry (1997), 62 (25), 8930-8931CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)A potential advantage of the pulsed field gradient diffusion NMR method over the chem. shift method is that binding is detected by the observation of the ligand NMR spectrum and not by changes in the receptor NMR spectrum. This should allow the expt. to be tuned to the binding affinity of the ligand by changing the relative receptor concn. This premise was tested by making a mixt. of four carboxylic acids and using the hydroquinine 9-phenanthryl ether as the receptor. The binding consts. for the ligands, and the pKa's are given. As the concn. of hydroquinine 9-phenanthryl ether is increased, the other carboxylic acids, sequentially begin to appear in their order of binding affinity. This demonstrates that the system can be tuned to a desired sensitivity level and that mixts. of compds. contg. only a few components can be screened to select components which bind at a desired level of affinity.
- 51Lin, M.; Shapiro, M. J.; Wareing, J. R. Diffusion-Edited NMR–Affinity NMR for Direct Observation of Molecular Interactions. J. Am. Chem. Soc. 1997, 119, 5249– 5250, DOI: 10.1021/ja963654+51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXjt1Churc%253D&md5=af2acda8d726760aa199a7d728f25e0eDiffusion-Edited NMR-Affinity NMR for Direct Observation of Molecular InteractionsLin, Mengfen; Shapiro, Michael J.; Wareing, James R.Journal of the American Chemical Society (1997), 119 (22), 5249-5250CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)A small mixt. of compds. can be selectively edited to find and identify components involved in mol. interactions using Pulse Field Gradient NMR techniques. The structure of the ligand can be deduced directly, without requiring sepn. of the mixt., using the LED and TOCSY (DECODES) pulse sequence.
- 52Bleicher, K.; Lin, M.; Shapiro, M. J.; Wareing, J. R. Diffusion Edited NMR: Screening Compound Mixtures by Affinity NMR to Detect Binding Ligands to Vancomycin. J. Org. Chem. 1998, 63, 8486– 8490, DOI: 10.1021/jo981736652https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXmslKjsLk%253D&md5=e9097b03f289b3b553a46cb9819b07c7Diffusion Edited NMR: Screening Compound Mixtures by Affinity NMR to Detect Binding Ligands to VancomycinBleicher, Konrad; Lin, Mengfen; Shapiro, Michael J.; Wareing, James R.Journal of Organic Chemistry (1998), 63 (23), 8486-8490CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)Affinity NMR can be used to produce an edited NMR spectrum that identifies ligands that bind to vancomycin from soln. mixts. contg. nonbinding mols. The Diffusion EnCODEd Spectroscopy (DECODES) expt. performed directly on the same sample can be used to det. the structure of the binding ligands without the need for a phys. sepn. step. The all-D amino acid tetrapeptides DDFA and DDFS, known ligands for vancomycin, were identified in the presence of eight nonbinding tetrapeptides. The bound-ligand signals in the two-dimensional DECODES spectrum are readily identified by comparison with the spectral patterns of the vancomycin cross-peaks in the 2D total correlation spectroscopy and correlation spectroscopy spectra. The screening of soln. mixts. of mols. for direct detection of mol. interactions and structural identification of the interacting ligands provides a powerful new tool to complement methods, such as affinity MS, which rely on the phys. sepn. of mixt. components to identify mol. interactions. The soln. mixts. of compds. for screening by affinity NMR could come from any source where the components are in similar relative amts., including synthesis by combinatorial chem. methods.
- 53Anderson, R. C.; Lin, M.; Shapiro, M. J. Affinity NMR: Decoding DNA Binding. J. Comb. Chem. 1999, 1, 69– 72, DOI: 10.1021/cc980004o53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXnsFKgt70%253D&md5=f8ec3e35bbe112639a0ec518a7be9a75Affinity NMR: Decoding DNA BindingAnderson, Robert C.; Lin, Mengfen; Shapiro, Michael J.Journal of Combinatorial Chemistry (1999), 1 (1), 69-72CODEN: JCCHFF; ISSN:1520-4766. (American Chemical Society)We have shown that affinity NMR can be used to edit a NMR spectrum so that ligands that have affinity to DNA can be obsd. in the presence of other nonbinding mols. Diffusion encoded spectroscopy (DECODES) can be used to identify the binding ligands. We were able to identify Hoechst 33342 as binding to the Drew-Dickerson dodecamer d(CGCGAATTCGCG)2 in the presence of the nonbinding mols. adenine, adenosine, and thiamine. Affinity NMR appears to be readily applicable to DNA systems for the following reasons. (1) The relaxation rate of the DNA oligonucleotides is favorable, thus the signal intensity loss due to relaxation is not severe. (2) A comparison of the patterns of the DNA cross-peaks upon binding in the two-dimensional total correlation spectroscopy and correlation spectroscopy spectrum are easily performed, and the ligand signals in the two-dimensional DECODES spectrum can be readily identified. (3) The arom. part of the DNA spectrum is devoid of 2D cross-peaks in these correlation spectra, greatly facilitating the interpretation of the bound ligand in the DECODES spectrum.
- 54Mayer, M.; Meyer, B. Characterization of ligand binding by saturation transfer difference NMR spectroscopy. Angew. Chem., Int. Ed. 1999, 38, 1784– 1788, DOI: 10.1002/(sici)1521-3773(19990614)38:12<1784::aid-anie1784>3.0.co;2-q54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXktVCntb0%253D&md5=03f825af40d22911062112dc6074cf46Characterization of ligand binding by saturation transfer difference NMR spectroscopyMayer, Moriz; Meyer, BerndAngewandte Chemie, International Edition (1999), 38 (12), 1784-1788CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH)The difference between a satn. transfer spectrum and a normal NMR spectrum provides a new and fast method (satn. transfer difference (STD) NMR spectroscopy) to screen compd. libraries for binding activity to proteins. STD NMR spectroscopy of mixts. of potential ligands with as little as 1 nmol of protein yields 1D and 2D NMR spectra that exclusively show signals from mols. with binding affinity. In addn., the ligand's binding epitope is easily identified because ligand residues in direct contact to the protein show much stronger STD signals. For example, the binding specificity of Lewisb-hexasaccharide to Aleulria aurantia agglutinin (AAA) can be mapped to the two fucosyl residues. The efficiency of the STD NMR technique is shown by the binding of N-acetylglucosamine (GlcNAc) to wheat germ agglutinin (WGA).
Supporting Information
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.8b05069.
Experimental procedures; spectral and computational details; and cartesian coordinates for all optimized structures (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.