Rose Bengal Binding to Collagen and Tissue Photobonding
- Emilio I. Alarcon
- ,
- Horacio Poblete
- ,
- HeeGwang Roh
- ,
- Jean-François Couture
- ,
- Jeffrey Comer
- , and
- Irene E. Kochevar
Abstract

We investigated two critical aspects of rose Bengal (RB) photosensitized protein cross-linking that may underlie recently developed medical applications. Our studies focused on the binding of RB to collagen by physical interaction and the effect of this binding and certain amino acids on RB photochemistry. Molecular dynamics simulations and free-energy calculation techniques, complemented with isothermal titration calorimetry, provided insight into the binding between RB and a collagen-like peptide (CLP) at the atomic level. Electrostatic interactions dominated, which is consistent with the finding that RB bound equally well to triple helical and single chain collagen. The binding free energy ranged from −5.7 to −3 kcal/mol and was strongest near the positively charged amino groups at the N-terminus and on lysine side chains. At high RB concentration, a maximum of 16 ± 3 bound dye molecules per peptide was found, which is consistent with spectroscopic evidence for aggregated RB bound to collagen or the CLP. Within a tissue-mimetic collagen matrix, RB photobleached rapidly, probably due to electron transfer to certain protein amino acids, as was demonstrated in solutions of free RB and arginine. In the presence of arginine and low oxygen concentrations, a product absorbing at 510 nm formed, presumably due to dehalogenation after electron transfer to RB. In the collagen matrix without arginine, the dye generated singlet oxygen as well as the 510 nm product. These results provide the first evidence of the effects of a tissue-like environment on the photochemical mechanisms of rose Bengal.
1 Introduction
2 Results
2.1 Association of Rose Bengal with Collagen and a Collagen-Like Peptide
2.1.1 RB–Collagen Association in Solution: Absorption Spectroscopy and Isothermal Titration Calorimetry
Figure 1

Figure 1. Effect of collagen-like peptide on 10 μM rose Bengal absorption. (a) Spectral changes and (b) changes in the absorption at RB monomer maximum 550 nm and at 570 nm, the maximum of the red-shifted peak for aggregated RB bound to CLP, plotted as a function of CLP concentration. All measurements were carried out in 10 mM MES buffer (pH 5.0).
2.1.2 Rose Bengal Association with Collagen Matrices
Figure 2

Figure 2. Right: Normalized absorption spectra for RB prepared at different dye concentrations incorporated within collagen matrices. Left: Normalized absorption spectra for RB measured in buffer solutions. All measurements were carried out in phosphate buffer pH 7.4 at room temperature.
2.1.3 Molecular Dynamics Simulations of RB–Collagen Association
Figure 3

Figure 3. Molecular simulation of RB binding to the CLP. (a) Molecular model of RB in the charge state used in the simulations. The carboxylate and phenolate moieties each have charges of −1. Atom color code: H, white; C, cyan; O, red; Cl, green; I, purple. (b) Snapshot of a simulation of the CLP triple helix with one molecule of RB. The CLP is shown as a green tube, Na+ and Cl– ions are shown as yellow and cyan spheres. For clarity, the explicit water molecules are shown as a transparent surface. (c) Free-energy landscape for RB in the vicinity of the CLP triple helix at low RB concentration, calculated by the adaptive biasing force method. The potential of mean force is mapped as a function of the position of the RB molecule along the CLP axis (disZ) and distance from this axis (disXY). The geometric contribution to the free energy along disXY has been removed. (d) Free-energy landscape at a higher RB concentration calculated from a set of equilibrium simulations including a CLP triple helix and 20 RB molecules. (e) Snapshot from the simulation with 20 RB molecules.
Figure 4

Figure 4. Molecular interactions between RB and CLP. (a) Electrostatic contact and H-bonding between the N-terminus of the CLP and the carboxylate group of RB. In these images, the phenolate (O–) group is highlighted by a red sphere to distinguish it from the carbonyl oxygen. Atoms are colored as in Figure 3, except that CLP carbons are colored dark green. H-bonds are indicated by dotted black lines. (b) Hydrophobic contact between the three-ring moiety of RB and a Pro residue of the CLP. (c) H-bonding between the O– group of RB and the backbone NH of the CLP, coupled with ionic contact between the O– group and a Lys side chain. (d) H-bonding between the OH group of hydroxyproline (Hyp) and the carboxylate group of RB. An ionic Lys–O– contact is also apparent. (e) The free-energy landscape of RB near the CLP with labels (a–d) indicating the location of RB in the corresponding panels of this figure. The ranges of disZ values occupied by the N-terminus and Asp, Hyp, and Lys side chains are also indicated. (f) The prevalence of H-bonding between RB and different groups of the CLP as a function of position along the CLP axis. Plotted are the number of H-bonds involving all CLP groups (CLP), backbone amide nitrogens (BB), N-terminal NH3+ groups (Nter), Hyp side chain OH groups, and Lys side chain NH3+ groups. (g) The prevalence of H-bonding between the CLP and different groups of RB as a function of position along the CLP axis.
2.2 Photochemical and Photophysical Behavior of Rose Bengal in Solution and 3D Collagen Matrix
2.2.1 Photodecomposition of Rose Bengal Alone in Solution
2.2.2 Influence of Arginine, Lysine, and Ascorbic Acid on Rose Bengal Photodegradation in Solution
Figure 5

Figure 5. Photodecomposition of rose Bengal (125 μM) in the presence of arginine (25 mM). (a) Irradiation of rose Bengal in nitrogen saturated solution. (b) Absorption spectra (expanded scale) of rose Bengal samples irradiated for 7 min in the presence of 25 mM Arg in solutions saturated with oxygen, air, and nitrogen. All data collected in PBS buffer pH 7.4 at room temperature. Solutions irradiated at 532 nm, irradiance = 0.3 W/cm2. Results shown are from single irradiations in triplicate independent experiments.
Figure 6

Figure 6. Photodegradation of 10 μM RB in the presence of different additives: (a) without and (b) with 2.5 μM collagen at RB/collagen (4:1). Left: Absorption spectra for RB measured in solution at different irradiation times. Right: Changes in absorption intensities measured at 550 nm, plotted as A/A0, for RB solutions recorded in the presence of different additives. All data collected in 10 mM MES buffer pH 5.0 at room temperature. Data correspond to the average calculated from three independent experiments each repeated in triplicate (n = 9).
2.2.2.1 Transient Absorption Measurements in Solution
2.2.3 Photodegradation of Rose Bengal in Collagen Solutions
2.2.4 Photodegradation of Rose Bengal in Collagen Hydrogel Matrices
Figure 7

Figure 7. Photodegradation of 40 μM RB incorporated within collagen hydrogels prepared using type I collagen containing different additives. Left: Absorption spectra for the RB–collagen hydrogel composite measured at different irradiation times in an air atmosphere. Right: Changes in absorption intensities measured at 556 nm for RB solutions recorded in the presence of different additives [Azide: 10 mM, Trp: 2 mM, Arg: 10 mM]. All data collected in PBS buffer pH 7.4 at room temperature. Data correspond to the average calculated from three independent experiments each repeated in triplicate (n = 9).
2.2.5 Transient Absorption Measurements in Collagen Matrices
Figure 8

Figure 8. Triplet transient lifetime and quenching of RB embedded in collagen hydrogels in the presence of different quenchers. (a) RB triplet decay monitored at 620 nm measured in air (black circles) or nitrogen (blue circles) saturated solutions. Insets correspond to decay residuals obtained from the exponential fit for the decays shown in the figure. Measured lifetimes for: (b) RB triplet, (c) RB anion radical, and (d) singlet oxygen under different conditions or additives. Concentrations of additives were: tryptophan: 2.0 mM, Arg: 10 mM, and sodium azide: 10 mM. Effect of additives for laser flash photolysis experiments were carried out in nitrogen saturated solutions. For singlet oxygen measurements, samples were equilibrated with air prior to laser excitation. Measurements for triplet and anion radical were carried out in 10 mM pH 5.0 MES buffer. For singlet oxygen phosphorescence, a 10 mM MES buffer was prepared with a pD of 5.0. In all cases, photodegradation was kept lower than 10%. Time traces correspond to the average of 12 separate decays from 3 independent samples.
2.2.6 Tryptophan-Mediated Photodegradation in Rose Bengal Containing Collagen Matrices
3 Discussion
4 Summary and Conclusions
5 Materials and Methods
5.1 Materials
5.2 Effect of Collagen and Collagen-Like Peptides on Rose Bengal Spectroscopic Properties
5.3 Changes in Collagen Surface Charge
5.4 Isothermal Titration Calorimetry (ITC)
5.5 Collagen Matrix Preparation and Characterization
5.6 Computer Simulations
5.6.1 Generation of Molecular Models
5.6.2 Classical Molecular Dynamics Simulation
5.6.3 Free-Energy Calculations
5.7 Light-Mediated Degradation of Rose Bengal in Homogenous Solution and 3D Collagen Matrix
5.8 Rose Bengal Laser Flash Photolysis
5.9 Statistical Analyses
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b00675.
Effect of type I porcine collagen, up to 2.5 μM, on the absorption spectra of 10 μM RB; representative circular dichroism spectra for a 1.25 μM rose Bengal solution with or without 5.0 μM type I collagen; isothermal titration calorimetry measurements for binding of rose Bengal to a collagen-like peptide (CLP); ζ potential measurements for 2.5 μM type I collagen in 10 mM pH 5.0 MES buffer in the presence of different rose Bengal concentrations; schematic representation for the preparation of RB containing collagen hydrogels; photodecomposition of rose Bengal (125 μM) under oxygen, air, and nitrogen saturated solutions; photodecomposition of RB (125 μM) in the presence of 1.25 mM sodium ascorbate in oxygen, air, and nitrogen saturated solutions; triplet transient lifetime and quenching of RB excited state in the presence of different quenchers; tryptophan degradation mediated by green light exposure in the RB containing collagen matrix (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.
Acknowledgment
The authors would like to thank Professor Tito Scaiano for the time-resolved measurements. E.I.A. thanks Michel Grenier from University of Ottawa for his assistance with the laser flash photolysis experiments. The authors would like also to express their gratitude to Dr. May Griffith from University of Montreal for providing a sample of the CLP peptide. I.E.K. and H.G.R. would like to thank Dr. Walfre Franco and Antonio Ortega-Martinez for technical support.
Arg | l-arginine |
CLP | collagen-like peptide |
PTB | photochemical tissue bonding |
RB | rose Bengal |
Trp | l-tryptophan |
References
This article references 51 other publications.
- 1Chan, B. P.; Amann, C.; Yaroslavsky, A. N.; Title, C.; Smink, D.; Zarins, B.; Kochevar, I. E.; Redmond, R. W. Photochemical repair of Achilles tendon rupture in a rat model J. Surg. Res. 2005, 124, 274– 279 DOI: 10.1016/j.jss.2004.09.019Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXjtVyrsrc%253D&md5=7021988a830cf8f0b6a4a21ef0a0f492Photochemical repair of Achilles tendon rupture in a rat modelChan, Barbara P.; Amann, Christopher; Yaroslavsky, Anna N.; Title, Craig; Smink, David; Zarins, Bertram; Kochevar, Irene E.; Redmond, Robert W.Journal of Surgical Research (2005), 124 (2), 274-279CODEN: JSGRA2; ISSN:0022-4804. (Elsevier)Photochem. tissue bonding (PTB) is an emerging technique for bonding or sealing tissue surfaces that requires light and a photoactive dye for its effect. The potential of PTB for tendon repair was assessed in a rat model. The optical properties of bovine tendon were detd. ex vivo to gauge the depth of light penetration as a function of wavelength and dosimetry parameters were established for PTB repair of ruptured tendon. PTB was then tested in vivo to repair transected tendons in Sprague-Dawley rats. Repair strengths were measured using a strain gauge up to 14 days post treatment. The effective penetration depth in tendon was estd. to be 0.68 mm at 514 nm. Following PTB treatment of mech. ruptured tendon, significant bonding was dependent on the presence of both light and dye and attained a plateau strength at a fluence of 125 J/cm2. In a subsequent in vivo study to investigate PTB for repair of transected rat Achilles tendon, the ultimate stress required to break the repaired tendon was measured immediately after irradn. and at 7 and 14 days post-repair. Results showed that the difference in the ultimate stress between control and PTB treatment groups was statistically significant immediately after treatment and at 7 days (p = 0.04) but not 14 days (p = 0.75) post-repair. PTB provides a benefit to tendon repair at early stages in repair and is worthy of further investigation as a potential surgical adjunct for tendon repair in orthopedic surgeries.
- 2Fairbairn, N. G.; Ng-Glazier, J.; Meppelink, A. M.; Randolph, M. A.; Valerio, I. L.; Fleming, M. E.; Kochevar, I. E.; Winograd, J. M.; Redmond, R. W. Light-Activated Sealing of Acellular Nerve Allografts following Nerve Gap Injury J. Reconstr. Microsurg. 2016, 32, 421– 430 DOI: 10.1055/s-0035-1571247Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28jgsleqtw%253D%253D&md5=0c21b1b08b0e14cfd2d3fb2fd3ad5ffcLight-Activated Sealing of Acellular Nerve Allografts following Nerve Gap InjuryFairbairn Neil G; Ng-Glazier Joanna; Meppelink Amanda M; Randolph Mark A; Winograd Jonathan M; Valerio Ian L; Fleming Mark E; Kochevar Irene E; Redmond Robert WJournal of reconstructive microsurgery (2016), 32 (6), 421-30 ISSN:.Introduction Photochemical tissue bonding (PTB) uses visible light to create sutureless, watertight bonds between two apposed tissue surfaces stained with photoactive dye. In phase 1 of this two-phase study, nerve gaps repaired with bonded isografts were superior to sutured isografts. When autograft demand exceeds supply, acellular nerve allograft (ANA) is an alternative although outcomes are typically inferior. This study assesses the efficacy of PTB when used with ANA. Methods Overall 20 male Lewis rats had 15-mm left sciatic nerve gaps repaired using ANA. ANAs were secured using epineurial suture (group 1) or PTB (group 2). Outcomes were assessed using sciatic function index (SFI), gastrocnemius muscle mass retention, and nerve histomorphometry. Historical controls from phase 1 were used to compare the performance of ANA with isograft. Statistical analysis was performed using analysis of variance and Bonferroni all-pairs comparison. Results All ANAs had signs of successful regeneration. Mean values for SFI, muscle mass retention, nerve fiber diameter, axon diameter, and myelin thickness were not significantly different between ANA + suture and ANA + PTB. On comparative analysis, ANA + suture performed significantly worse than isograft + suture from phase 1. However, ANA + PTB was statistically comparable to isograft + suture, the current standard of care. Conclusion Previously reported advantages of PTB versus suture appear to be reduced when applied to ANA. The lack of Schwann cells and neurotrophic factors may be responsible. PTB may improve ANA performance to an extent, where they are equivalent to autograft. This may have important clinical implications when injuries preclude the use of autograft.
- 3Mulroy, L.; Kim, J.; Wu, I.; Scharper, P.; Melki, S. A.; Azar, D. T.; Redmond, R. W.; Kochevar, I. E. Photochemical keratodesmos for repair of lamellar corneal incisions Invest. Ophthalmol. Visual Sci. 2000, 41, 3335– 3340Google ScholarThere is no corresponding record for this reference.
- 4O’Neill, A. C.; Winograd, J. M.; Zeballos, J. L.; Johnson, T. S.; Randolph, M. A.; Bujold, K. E.; Kochevar, I. E.; Redmond, R. W. Microvascular anastomosis using a photochemical tissue bonding technique Lasers Surg. Med. 2007, 39, 716– 722 DOI: 10.1002/lsm.20548Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BD2snkvVehsQ%253D%253D&md5=0d0b5e9d0759a9e621613c966539ded3Microvascular anastomosis using a photochemical tissue bonding techniqueO'Neill Anne C; Winograd Jonathan M; Zeballos Jose L; Johnson T Shane; Randolph Mark A; Bujold Kenneth E; Kochevar Irene E; Redmond Robert WLasers in surgery and medicine (2007), 39 (9), 716-22 ISSN:0196-8092.BACKGROUND AND OBJECTIVES: Photochemical tissue bonding (PTB) combines photoactive dyes with visible light to create fluid-tight seals between tissue surfaces without causing collateral thermal damage. The potential of PTB to improve outcomes over standard of care microsurgical reanastomoses of blood vessels in ex vivo and in vivo models was evaluated. STUDY DESIGN: The mechanical strength and integrity of PTB and standard microsurgical suture repairs in ex vivo porcine brachial arteries (n = 10) were compared using hydrostatic testing of leak point pressure (LPP). Femoral artery repair in vivo was measured in Sprague-Dawley rats using either standard microvascular sutures (n = 20) or PTB (n = 20). Patency was evaluated at 6 hours (n = 10) and 8 weeks post-repair (n = 10) for each group. RESULTS: PTB produced significantly higher LPPs (1,100+/- 150 mmHg) than suture repair (350+/-40 mmHg, P<0.001) in an ex vivo study. In an in vivo study all femoral arteries in both suture and PTB repair groups were patent at 6 hours post-repair. At 8 weeks post-repair the patency rate was 80% for both groups. No evidence of aneurysm formation was seen in either group and bleeding was absent from the repair site in the PTB-treated vessels, in contrast to the suture repair group. CONCLUSION: PTB is a feasible microvascular repair technique that results in an immediate, mechanically robust bond with short- and long-term patency rates equal to those for standard suture repair.
- 5Senthil-Kumar, P.; Ni, T.; Randolph, M. A.; Velmahos, G. C.; Kochevar, I. E.; Redmond, R. W. A light-activated amnion wrap strengthens colonic anastomosis and reduces peri-anastomotic adhesions Lasers Surg. Med. 2016, 48, 530– 537 DOI: 10.1002/lsm.22507Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28fitVWgug%253D%253D&md5=c2ef2ce184a37b1f44bfbafe4dd08bb8A light-activated amnion wrap strengthens colonic anastomosis and reduces peri-anastomotic adhesionsSenthil-Kumar Prabhu; Ni Tao; Kochevar Irene E; Redmond Robert W; Senthil-Kumar Prabhu; Randolph Mark A; Ni Tao; Velmahos George CLasers in surgery and medicine (2016), 48 (5), 530-7 ISSN:.BACKGROUND AND OBJECTIVE: Colonic anastomotic failure is a dreaded complication, and multiple surgical techniques have failed to eliminate it. Photochemical tissue bonding (PTB) is a method of sealing tissue surfaces by light-activated crosslinking. We evaluated if a human amniotic membrane (HAM), sealed over the anastomotic line by PTB, increases the anastomotic strength. STUDY DESIGN: Sprague-Dawley rats underwent midline laparotomy followed by surgical transection of the left colon. Animals were randomized to colonic anastomosis by one of the following methods (20 per group): single-layer continuous circumferential suture repair (SR); SR with a HAM wrap attached by suture (SR+ HAM-S); SR with HAM bonded photochemically over the anastomotic site using 532 nm light (SR+ HAM-PTB); approximation of the bowel ends with only three sutures and sealing with HAM-PTB (3+ HAM-PTB). A control group underwent laparotomy alone with no colon resection (NR). Sub-groups (n = 10) were sacrificed at days 3 and 7 post-operatively and adhesions were evaluated. A 6 cm section of colon was then removed and strength of anastomosis evaluated by burst pressure (BP) measurement. RESULTS: A fourfold increase in BP was observed in the SR+ HAM-PTB group compared to suture repair alone (94 ± 3 vs. 25 ± 8 mm Hg, P < 0.0001) at day 3. At day 7 the burst pressures were 165 ± 40 and 145 ± 31 mm Hg (P = 1), respectively. A significant decrease in peri-anastomotic adhesions was observed in the SR+ HAM-PTB group compared to the SR group at both time points (P < 0.001). CONCLUSION: Sealing sutured colonic anastomotic lines with HAM-PTB increases the early strength of the repair and reduces peri-anastomotic adhesions. Lasers Surg. Med. 48:530-537, 2016. © 2016 Wiley Periodicals, Inc.
- 6Tsao, S.; Yao, M.; Tsao, H.; Henry, F. P.; Zhao, Y.; Kochevar, J. J.; Redmond, R. W.; Kochevar, I. E. Light-activated tissue bonding for excisional wound closure: a split-lesion clinical trial Br. J. Dermatol. 2012, 166, 555– 563 DOI: 10.1111/j.1365-2133.2011.10710.xGoogle Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC383lsVCgug%253D%253D&md5=26e81b4b36431acc477589ade5ec3d80Light-activated tissue bonding for excisional wound closure: a split-lesion clinical trialTsao S; Yao M; Tsao H; Henry F P; Zhao Y; Kochevar J J; Redmond R W; Kochevar I EThe British journal of dermatology (2012), 166 (3), 555-63 ISSN:.BACKGROUND: Apposition of wound edges by sutures provides a temporary scaffold and tension support for healing. We have developed a novel tissue-sealing technology, photoactivated tissue bonding (PTB), which immediately crosslinks proteins between tissue planes, thereby sealing on a molecular scale. OBJECTIVES: To determine the effectiveness of PTB for superficial closure of skin excisions and to compare the results with standard epidermal suturing. METHODS: A split-lesion, paired comparison study of 31 skin excisions was performed. Following deep closure with absorbable sutures, one-half of each wound was superficially closed with nonabsorbable nylon sutures while the other half was stained with Rose Bengal dye and treated with green light. Overall appearance and scar characteristics were rated at 2weeks and 6months in a blinded manner by three dermatologists viewing photographs, by two onsite physicians and by patients. RESULTS: At 2weeks, neither sutured nor PTB-treated segments showed dehiscence; however, PTB-sealed segments showed less erythema than sutured segments as determined by photographic (P=0·001) and onsite evaluations (P=0·005). Overall appearance after PTB was judged better than after sutures (P=0·002). At 6months, scars produced by PTB were deemed superior to scars resulting from sutures in terms of appearance (P<0·001), width (P=0·002) and healing (P=0·003). Patients were more satisfied with the appearance of the PTB-sealed wound half after 2weeks and 6months (P=0·013 and P=0·003, respectively). CONCLUSIONS: A novel molecular suturing technique produces effective wound sealing and less scarring than closure with nylon interrupted epidermal sutures. Comparisons with better suturing techniques are warranted.
- 7Verter, E. E.; Gisel, T. E.; Yang, P.; Johnson, A. J.; Redmond, R. W.; Kochevar, I. E. Light-initiated bonding of amniotic membrane to cornea Invest. Ophthalmol. Visual Sci. 2011, 52, 9470– 9477 DOI: 10.1167/iovs.11-7248Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjtVKkurY%253D&md5=2055c585b721cfd26530c542e03fd7a9Light-initiated bonding of amniotic membrane to corneaVerter, E. Eri; Gisel, Thomas E.; Yang, Penggao; Johnson, Anthony J.; Redmond, Robert W.; Kochevar, Irene E.Investigative Ophthalmology & Visual Science (2011), 52 (13), 9470-9477CODEN: IOVSDA; ISSN:1552-5783. (Association for Research in Vision and Ophthalmology)Purpose: Suturing amniotic membrane to cornea during surgery is time consuming, and sutures may further damage the eye. The authors introduce a novel sutureless, light-activated technique that securely attaches amnion to cornea through protein-protein crosslinks. Methods: Cryopreserved human amniotic membrane, stained with Rose Bengal (RB), was placed over a full-thickness wound in deepithelialized rabbit cornea and was treated with green laser. The intraocular pressure that broke the seal (IOPL) was measured, and adhesion was measured with a peel test. The influences on bonding strength of fluence, irradiance, RB concn., and amnion surface bonded were measured. Epithelial cell migration on treated amnion and keratocyte viability after bonding were also measured. The involvement in the bonding mechanism of oxygen, singlet oxygen, and assocn. of RB with stromal collagen was investigated. Results: Sealing amniotic membrane over cornea using 0.1% RB and 150 J/cm2 at 532 nm produced an IOPL of 261 ± 77 mm Hg ex vivo and 448 mm ± 212 mm Hg in vivo. The ex vivo IOPL increased with increasing fluence (50-150 J/cm2). Equivalent IOPL was produced for bonding basement membrane or stromal amnion surfaces. The bonding treatment was not toxic to keratocytes but slightly reduced the migration of corneal epithelial cells on amnion ex vivo. Mechanism studies indicated that RB forms two complexes with amnion stromal collagen, that bonding requires oxygen, and that singlet oxygen mediates protein crosslinking. Conclusions: A rapid, light-activated technique produces strong, immediate bonding between amnion and cornea and merits further evaluation for ocular surface surgeries.
- 8Cherfan, D.; Verter, E. E.; Melki, S.; Gisel, T. E.; Doyle, F. J., Jr.; Scarcelli, G.; Yun, S. H.; Redmond, R. W.; Kochevar, I. E. Collagen cross-linking using rose bengal and green light to increase corneal stiffness Invest. Ophthalmol. Visual Sci. 2013, 54, 3426– 3433 DOI: 10.1167/iovs.12-11509Google ScholarThere is no corresponding record for this reference.
- 9Zhu, H.; Alt, C.; Webb, R. H.; Melki, S.; Kochevar, I. E. Corneal Crosslinking With Rose Bengal and Green Light: Efficacy and Safety Evaluation Cornea 2016, 35, 1234– 1241 DOI: 10.1097/ICO.0000000000000916Google ScholarThere is no corresponding record for this reference.
- 10Goldstone, R. N.; McCormack, M. C.; Goldstein, R. L.; Mallidi, S.; Randolph, M. A.; Watkins, M. T.; Redmond, R. W.; Austen, W. G., Jr. Photochemical Tissue Passivation Attenuates AV Fistula Intimal HyperplasiaAnn. Surg. 2016, DOI: 10.1097/SLA.0000000000002046 .Google ScholarThere is no corresponding record for this reference.
- 11Goldstone, R. N.; McCormack, M. C.; Khan, S. I.; Salinas, H. M.; Meppelink, A.; Randolph, M. A.; Watkins, M. T.; Redmond, R. W.; Austen, W. G., Jr. Photochemical Tissue Passivation Reduces Vein Graft Intimal Hyperplasia in a Swine Model of Arteriovenous Bypass Grafting J. Am. Heart Assoc. 2016, 5e003856 DOI: 10.1161/JAHA.116.003856Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvF2jsrjN&md5=55d0373eea312e7f41d626f8a28ecf39Photochemical tissue passivation reduces vein graft intimal hyperplasia in a swine model of arteriovenous bypass graftingGoldstone, Robert N.; McCormack, Michael C.; Khan, Saiqa I.; Salinas, Harry M.; Meppelink, Amanda; Randolph, Mark A.; Watkins, Michael T.; Redmond, Robert W.; Austen, William G. Jr.Journal of the American Heart Association (2016), 5 (8), e003856/1-e003856/10CODEN: JAHABZ; ISSN:2047-9980. (Wiley-Blackwell)Background-Bypass grafting remains the std. of care for coronary artery disease and severe lower extremity ischemia. Efficacy is limited by poor long-term venous graft patency secondary to intimal hyperplasia (IH) caused by venous injury upon exposure to arterial pressure. We investigate whether photochem. tissue passivation (PTP) treatment of vein grafts modulates smooth muscle cell (SMC) proliferation and migration, and inhibits development of IH. Methods and Results-PTP was performed at increasing fluences up to 120 J/cm2 on porcine veins. Tensiometry performed to assess vessel elasticity/stiffness showed increased stiffness with increasing fluence until plateauing at 90 J/cm2 (median, interquartile range [IQR]). At 90 J/cm2, PTP-treated vessels had a 10-fold greater Young's modulus than untreated controls (954 [IQR, 2217] vs 99 kPa [IQR, 63]; P=0.03). Each pig received a PTP-treated and untreated carotid artery venous interposition graft. At 4-wk, intimal/medial areas were assessed. PTP reduced the degree of IH by 66% and medial hypertrophy by 49%. Intimal area was 3.91 (IQR, 1.2) and 1.3 mm2 (IQR, 0.97; P≤0.001) in untreated and PTP-treated grafts, resp. Medial area was 9.2 (IQR, 3.2) and 4.7 mm2 (IQR, 2.0; P≤0.001) in untreated and PTP-treated grafts, resp. Immunohistochem. was performed to assess alpha-smooth muscle actin (SMA) and proliferating cell nuclear antigen (PCNA). Objectively, there were less SMA-pos. cells within the intima/media of PTP-treated vessels than controls. There was an increase in PCNA-pos. cells within control vein grafts (18% [IQR, 5.3]) vs. PTP-treated vein grafts (5% [IQR, 0.9]; P=0.02). Conclusions-By strengthening vein grafts, PTP decreases SMC proliferation and migration, thereby reducing IH.
- 12Fernandes, J. R.; Salinas, H. M.; Broelsch, G. F.; McCormack, M. C.; Meppelink, A. M.; Randolph, M. A.; Redmond, R. W.; Austen, W. G., Jr. Prevention of capsular contracture with photochemical tissue passivation Plast. Reconstr. Surg. 2014, 133, 571– 577 DOI: 10.1097/01.prs.0000438063.31043.79Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXjt1SntL0%253D&md5=a423afa38ba1d0cf77a9dab0853efb4dPrevention of Capsular Contracture with Photochemical Tissue PassivationFernandes, Justin R.; Salinas, Harry M.; Broelsch, G. Felix; McCormack, Michael C.; Meppelink, Amanda M.; Randolph, Mark A.; Redmond, Robert W.; Austen, William G., Jr.Plastic and Reconstructive Surgery (2014), 133 (3), 571-577CODEN: PRSUAS; ISSN:1529-4242. (Lippincott Williams & Wilkins)Background: Capsular contracture is the most common complication following the insertion of breast implants. Within a decade, half of patients will develop capsular contracture, leading to significant morbidity and need for reoperation. There is no preventative treatment available and the recurrence rate remains high. Photochem. tissue passivation is a novel tissue-stabilization technique that results in collagen crosslinking. It can rapidly link collagen fibers in situ, preserving normal tissue architecture. By using this therapy to passivate the collagenous tissues of the implant pocket, the authors hope to prevent the development of pathogenic collagen bundles and subsequent capsule contracture. Methods: Six-cubic centimeter tissue expanders were placed below the panniculus carnosus muscle along the dorsum of New Zealand white rabbits. Fibrin glue was instilled into each implant pocket to induce contracture. Treated pockets received photochem. tissue passivation by coating them with a photosensitizing dye and exposing the area to a 532-nm light. After 8 wk, capsule tissue was harvested for histol. evaluation. Results: Implant capsule thickness is the no. one prognostic factor for contracture development. The authors demonstrated a 52 percent decrease in capsule thickness in the passivated group compared with controls. Photochem. tissue passivation resulted in fewer fibrohistiocytic cells and macrophages and in reduced synovial metaplasia and smooth muscle actin deposition. Conclusions: Photochem. tissue passivation significantly decreased both capsule thickness and smooth muscle actin deposition. It is a promising technique for preventing capsular contracture that can be performed at the time of initial surgery without a significant increase in procedure time.
- 13Ludvikova, L.; Fris, P.; Heger, D.; Sebej, P.; Wirz, J.; Klan, P. Photochemistry of rose bengal in water and acetonitrile: a comprehensive kinetic analysis Phys. Chem. Chem. Phys. 2016, 18, 16266– 16273 DOI: 10.1039/C6CP01710JGoogle Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XptVCrurY%253D&md5=16827a5eff05f8af311b79d04e80a169Photochemistry of rose bengal in water and acetonitrile: a comprehensive kinetic analysisLudvikova, Lucie; Fris, Pavel; Heger, Dominik; Sebej, Peter; Wirz, Jakob; Klan, PetrPhysical Chemistry Chemical Physics (2016), 18 (24), 16266-16273CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The photophys. and photochem. properties of rose bengal (RB) in degassed aq. and acetonitrile solns. were studied using steady-state and transient absorption spectroscopies. This comprehensive investigation provides detailed information about the kinetics and the optical properties of all intermediates involved: the triplet excited state and the oxidized and reduced forms of RB. A full kinetic description is used to control the concns. of these intermediates by changing the initial exptl. conditions.
- 14Neckers, D. C. Rose Bengal J. Photochem. Photobiol. A 1989, 47, 1– 29 DOI: 10.1016/1010-6030(89)85002-6Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXktlWisbw%253D&md5=532dc3c2c508b43964c9247be4264613Rose BengalNeckers, D. C.Journal of Photochemistry and Photobiology, A: Chemistry (1989), 47 (1), 1-29CODEN: JPPCEJ; ISSN:1010-6030.A review with 38 refs. The spectral properties, photochem. reactivity, and photophys. parameters of all of the known derivs. of Rose Bengal, 2,4,5,7-tetraiodo-3',4',5',6'-tetrachlorofluorescein, are reported and compared.
- 15Fleming, G. R.; Morris, J. M.; Morrison, R. J. S.; Robinson, G. W. Picosecond fluorescence studies of xanthene dyes J. Am. Chem. Soc. 1977, 99, 4306– 4311 DOI: 10.1021/ja00455a017Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2sXkvVCjsrk%253D&md5=d2ae45a7fa0caf469a6dcf0b5cc9a0e1Picosecond fluorescence studies of xanthene dyesFleming, G. R.; Knight, A. W. E.; Morris, J. M.; Morrison, R. J. S.; Robinson, G. W.Journal of the American Chemical Society (1977), 99 (13), 4306-11CODEN: JACSAT; ISSN:0002-7863.Subnanosecond lifetime measurements using picosecond pulses from a mode-locked Nd3+-glass laser together with conventional absorption and fluorescence-yield methods are used to study the photophys. of fluorescein, eosin, erythrosin, and rose bengal in aq. and alc. solvents. For each of the dye mols., absorption and fluorescence max. move towards higher energy (blue shift) as the solvent changes from Me2CHOH to H2O. Fluorescence lifetimes and quantum yields decrease with this solvent change and also with increased halogenation of the fluorescein parent, owing to the heavy-atom effect. The variations obsd. in the nonradiative part of the decay rate are attributed to variations in the rate of S1-T1 intersystem crossing. For these particular solvent-solute combinations stabilization energies lie in the order ΔE(T1) < ΔE(S1) < ΔE(S0). This is consistent with both the increased S1-S0 spectral blue shifts and the enhanced intersystem crossing rate, arising from a smaller S1-T1 energy gap, when these dye mols. are placed in a more aq. solvent environment. The use of these dyes as fluorescent probes in biol. important mols. is discussed.
- 16Kochevar, I. E.; Redmond, R. W. Photosensitized production of singlet oxygen Methods Enzymol. 2000, 319, 20– 28 DOI: 10.1016/s0076-6879(00)19004-4Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXlsl2lsbo%253D&md5=46bb84b5807a1fe38a6afa20fc977996Photosensitized production of singlet oxygenKochevar, Irene E.; Redmond, Robert W.Methods in Enzymology (2000), 319 (), 20-28CODEN: MENZAU; ISSN:0076-6879. (Academic Press)A review with 8 refs. Photosensitization is a simple and controllable method for the generation of singlet oxygen in soln. and in cells. Methods are described for detg. the yield of singlet oxygen in soln., for measurement of the rate of reaction between single oxygen and a substrate, and for comparing the effectiveness of singlet oxygen generated by different photosensitizers in cells. These quant. measurements can lead to better understanding of the interaction of singlet oxygen with biomols. (c) 2000 Academic Press.
- 17Shen, H. R.; Spikes, J. D.; Kopeckova, P.; Kopecek, J. Photodynamic crosslinking of proteins. II. Photocrosslinking of a model protein-ribonuclease A J. Photochem. Photobiol. B 1996, 35, 213– 219 DOI: 10.1016/S1011-1344(96)07300-9Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xms1emurk%253D&md5=b2f68b4acb7c761ec67244bdf0eab706Photodynamic crosslinking of proteins. II. Photocrosslinking of a model protein-ribonuclease AShen, Hui-Rong; Spikes, John D.; Kopeckova, Pavla; Kopecek, JindrichJournal of Photochemistry and Photobiology, B: Biology (1996), 35 (3), 213-219CODEN: JPPBEG; ISSN:1011-1344. (Elsevier)Illumination of bovine pancreatic RNase A (RNase A) in soln. in the presence of rose bengal as a photosensitizer resulted in the progressive formation of enzyme dimers, trimers, tetramers and higher oligomers, as measured by gel electrophoresis and size exclusion chromatog. Oxygen was necessary for crosslink formation, and azide inhibition studies indicated that singlet oxygen was involved in the process. Chem. modification of His residues (with di-Et pyrocarbonate) and/or Lys residues (with acetic acid N-hydroxysuccinimide ester) in the enzyme decreased crosslinking, suggesting the participation of these two amino acid residues in the reaction. Met and cystine residues did not appear to be involved. Similar studies have shown that model N-(2-hydroxypropyl)methacrylamide (HPMA) copolymers contg. ε-aminocaproic acid side chains terminating in His or Lys residues are photodynamically crosslinked via His-His or His-Lys interactions. Treatment of crosslinked RNase A and its His, Lys and Lys-His derivs. for 5 min at 97 °C in a dithiothreitol-sodium dodecyl sulfate mixt. efficiently ruptured a major part of the photodynamically formed crosslinks; treatment with the detergent alone had no effect. Similar results were previously obtained with the crosslinked amino acid-contg. HPMA copolymers, suggesting that photodynamic crosslinks involving His-His and His-Lys interaction are chem. the same in RNase A and the copolymer model.
- 18Lambert, C. R.; Kochevar, I. E. Electron transfer quenching of the rose bengal triplet state Photochem. Photobiol. 1997, 66, 15– 25 DOI: 10.1111/j.1751-1097.1997.tb03133.xGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXksF2ksbw%253D&md5=d9f0db65a5fe55e35ea986835222e958Electron transfer quenching of the rose bengal triplet stateLambert, Christopher R.; Kochevar, Irene E.Photochemistry and Photobiology (1997), 66 (1), 15-25CODEN: PHCBAP; ISSN:0031-8655. (American Society for Photobiology)The potential for electron transfer quenching of rose bengal triplet (3RB2-) to compete with energy transfer quenching by oxygen was evaluated. Rate consts. for oxidative and reductive quenching were measured in buffered aq. soln., acetonitrile and in small unilamellar liposomes using laser flash photolysis. Biol. relevant quenchers were used that varied widely in structure, redn. potential and charge. Radical ion yields (φi) were measured by monitoring the absorption of the rose bengal semireduced (RB·3-) and semioxidized (RB·-) radicals. The results in soln. were analyzed as a function of the free energy for electron transfer (ΔG) calcd. using the Weller equation including electrostatic terms. Exothermic oxidative quenching was about 10-fold faster than exothermic reductive quenching in aq. soln. The quenching rate consts. decreased as ΔG approached zero in both aq. and acetonitrile soln. Exceptions to these generalizations were obsd. that could be rationalized by specific steric or electrostatic effects or by a change in mechanism. The results suggest that electron transfer reactions with some potential quenchers in cells could compete with formation of singlet oxygen [O2(1Δg)]. Values of φi were generally greater for reductive quenching and, for oxidative quenching, greater in acetonitrile than in buffer. Electron transfer quenching of 3RB2- in liposomes, below the phase transition temp. was slower than in soln. for both lipid-sol. and water-sol. quenchers indicating that these reactions may not compete with formation of (O21Δg) during cell photosensitization.
- 19Zakrzewski, A.; Neckers, D. C. Bleaching products of rose bengal under reducing conditions Tetrahedron 1987, 43, 4507– 4512 DOI: 10.1016/S0040-4020(01)86891-5Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXkslaku70%253D&md5=e1bb13110d942833b5df2b176055e93eBleaching products of Rose Bengal under reducing conditionsZakrzewski, Andrzej; Neckers, D. C.Tetrahedron (1987), 43 (20), 4507-12CODEN: TETRAB; ISSN:0040-4020.The bleaching behavior of Rose Bengal under reducing conditions was elucidated by detg. the products of chem. and photochem. redn. of Rose Bengal.
- 20Oster, G.; Oster, G. K.; Karg, G. Extremely long-lived intermediates in photochemical reactions of dyes in non-visous media J. Phys. Chem. 1962, 66, 2514– 2517 DOI: 10.1021/j100818a045Google ScholarThere is no corresponding record for this reference.
- 21Luttrull, D. K.; Valdes-Aguilera, O.; Linden, S. M.; Paczkowski, J.; Neckers, D. C. Rose Bengal aggregation in rationally synthesized dimeric systems Photochem. Photobiol. 1988, 47, 551– 557 DOI: 10.1111/j.1751-1097.1988.tb08843.xGoogle ScholarThere is no corresponding record for this reference.
- 22Simpson, M. J.; Poblete, H.; Griffith, M.; Alarcon, E. I.; Scaiano, J. C. Impact of dye-protein interaction and silver nanoparticles on rose bengal photophysical behavior and protein photocrosslinking Photochem. Photobiol. 2013, 89, 1433– 1441 DOI: 10.1111/php.12119Google ScholarThere is no corresponding record for this reference.
- 23Valdes-Aguilera, O.; Neckers, D. C. Aggregation phenomena in xanthene dyes Acc. Chem. Res. 1989, 22, 171– 177 DOI: 10.1021/ar00161a002Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXitFCksb8%253D&md5=2feeb8679bb8c7cf735a59e3aa19590aAggregation phenomena in xanthene dyesValdes-Aguilera, Oscar; Neckers, D. C.Accounts of Chemical Research (1989), 22 (5), 171-7CODEN: ACHRE4; ISSN:0001-4842.A crit. review with 63 refs. on aggregation of xanthene dyes, principally rhodamines and fluoresceins. The author's work on Rose Bengal is considered in detail.
- 24Xu, D.; Neckers, D. C. Aggregation of rose bengal molecules in solution J. Photochem. Photobiol. A 1987, 40, 361– 370 DOI: 10.1016/1010-6030(87)85013-XGoogle Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXmtlymtL0%253D&md5=239ea45bfc0643e238061f40c67e6130Aggregation of Rose Bengal molecules in solutionXu, Danian; Neckers, D. C.Journal of Photochemistry and Photobiology, A: Chemistry (1987), 40 (2-3), 361-70CODEN: JPPCEJ; ISSN:1010-6030.Absorption, emission, and excitation spectral data support the hypothesis that Rose Bengal forms H-type aggregates in water and in polar, protic solvents.
- 25Alarcón, E.; Edwards, A. M.; Aspee, A.; Borsarelli, C. D.; Lissi, E. A. Photophysics and photochemistry of rose bengal bound to human serum albumin Photochem. Photobiol. Sci. 2009, 8, 933– 943 DOI: 10.1039/b901056dGoogle Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXotVenu7s%253D&md5=c520abf9b3cdc4448c16998478eafef5Photophysics and photochemistry of rose bengal bound to human serum albuminAlarcon, Emilio; Edwards, Ana Maria; Aspee, Alexis; Borsarelli, Claudio D.; Lissi, Eduardo A.Photochemical & Photobiological Sciences (2009), 8 (7), 933-943CODEN: PPSHCB; ISSN:1474-905X. (Royal Society of Chemistry)Rose bengal (RB) readily binds to human serum albumin (HSA). At low RB concns., 90% of the dye is assocd. to the protein (5 μM). This assocn. takes place in specific binding sites I and/or II. At higher RB concns., unspecific binding takes place with up to 10 RB mols. bound per protein mol. The behavior of excited RB mols. bound to HSA is widely different to that obsd. in aq. soln. Furthermore, the data also show that the behavior of bound RB mols. changes with the av. no. of dye mols. per protein (n). In particular, when n is large, the fluorescence yield is significantly reduced and no measurable long-lived triples and free singlet oxygen formation from bound dyes is detected. These results are related to self-quenching of the singlet and, most likely, excited triplets. All results point to the relevance of intra-protein generated singlet oxygen. However, when the dye is bound to the protein, at low oxygen concns. such as those prevailing in vivo, trapping by oxygen of the triplet becomes inefficient and type I processes could contribute to the obsd. photoprocesses.
- 26Ni, T.; Senthil-Kumar, P.; Dubbin, K.; Aznar-Cervantes, S. D.; Datta, N.; Randolph, M. A.; Cenis, J. L.; Rutledge, G. C.; Kochevar, I. E.; Redmond, R. W. A photoactivated nanofiber graft material for augmented Achilles tendon repair Lasers Surg. Med. 2012, 44, 645– 652 DOI: 10.1002/lsm.22066Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC38bgt1WhtQ%253D%253D&md5=f59c0a45f8df812d81e675aa7e0a2011A photoactivated nanofiber graft material for augmented Achilles tendon repairNi Tao; Senthil-Kumar Prabhu; Dubbin Karen; Aznar-Cervantes Salvador D; Datta Neha; Randolph Mark A; Cenis Jose L; Rutledge Gregory C; Kochevar Irene E; Redmond Robert WLasers in surgery and medicine (2012), 44 (8), 645-52 ISSN:.BACKGROUND AND OBJECTIVE: Suture repair of Achilles tendon rupture can cause infection, inflammation and scarring, while prolonged immobilization promotes adhesions to surrounding tissues and joint stiffness. Early mobilization can reduce complications provided the repair is strong enough to resist re-rupture. We have developed a biocompatible, photoactivated tendon wrap from electrospun silk (ES) to provide additional strength to the repair that could permit early mobilization, and act as a barrier to adhesion formation. STUDY DESIGN/MATERIAL AND METHODS: ES nanofiber mats were prepared by electrospinning. New Zealand white rabbits underwent surgical transection of the Achilles tendon and repair by: (a) SR: standard Kessler suture + epitendinous suture (5-0 vicryl). (b) ES/PTB: a single stay suture and a section of ES mat, stained with 0.1% Rose Bengal (RB), wrapped around the tendon and bonded with 532 nm light (0.3 W/cm(2) , 125 J/cm(2) ). (c) SR + ES/PTB: a combination of (a) and (b). Gross appearance, extent of adhesion formation and biomechanical properties of the repaired tendon were evaluated at Days 7, 14, or 28 post-operatively (n = 8 per group at each time point). RESULTS: Ultimate stress (US) and Young's modulus (E) in the SR group were not significantly different from the ES/PTB group at Days 7 (US, P = 0.85; E, P = 1), 14 (US, P = 0.054; E, P = 1), and 28 (US, P = 0.198; E, P = 0.12) post-operatively. Adhesions were considerably greater in the SR group compared to the ES/PTB group at Days 7 (P = 0.002), 14 (P < 0.0001), and 28 (P < 0.0001). The combination approach of SR + ES/PTB gave the best outcomes in terms of E at 7 (P < 0.016) and 14 days (P < 0.016) and reduced adhesions compared to SR at 7 (P < 0.0001) and 14 days (P < 0.0001), the latter suggesting a barrier function for the photobonded ES wrap. CONCLUSION: Photochemical sealing of a ES mat around the tendon repair site provides considerable benefit in Achilles tendon repair. Lasers Surg. Med. 44: 645-652, 2012. © 2012 Wiley Periodicals, Inc.
- 27Stetefeld, J.; Frank, S.; Jenny, M.; Schulthess, T.; Kammerer, R. A.; Boudko, S.; Landwehr, R.; Okuyama, K.; Engel, J. Collagen stabilization at atomic level: crystal structure of designed (GlyProPro)10foldon Structure 2003, 11, 339– 346 DOI: 10.1016/S0969-2126(03)00025-XGoogle Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXhvFait7g%253D&md5=75307e701596341cc24da4e3ae1bd7c1Collagen Stabilization at Atomic Level Crystal Structure of Designed (GlyProPro)10foldonStetefeld, Jorg; Frank, Sabine; Jenny, Margrit; Schulthess, Therese; Kammerer, Richard A.; Boudko, Sergei; Landwehr, Ruth; Okuyama, Kenji; Engel, JurgenStructure (Cambridge, MA, United States) (2003), 11 (3), 339-346CODEN: STRUE6; ISSN:0969-2126. (Cell Press)In a designed fusion protein the trimeric domain foldon from bacteriophage T4 fibritin was connected to the C terminus of the collagen model peptide (GlyProPro)10 by a short Gly-Ser linker to facilitate formation of the three-stranded collagen triple helix. Crystal structure anal. at 2.6 A resoln. revealed conformational changes within the interface of both domains compared with the structure of the isolated mols. A striking feature is an angle of 62.5° between the symmetry axis of the foldon trimer and the axis of the triple helix. The melting temp. of (GlyProPro)10 in the designed fusion protein (GlyProPro)10foldon is higher than that of isolated (GlyProPro)10, which suggests an entropic stabilization compensating for the destabilization at the interface.
- 28Mirenda, M.; Dicelio, L. E.; San Roman, E. Effect of molecular interactions on the photophysics of Rose Bengal in polyelectrolyte solutions and self-assembled thin films J. Phys. Chem. B 2008, 112, 12201– 12207 DOI: 10.1021/jp803892gGoogle Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhtVymtLrK&md5=c6858f9042e0b1abc53c70d13981c743Effect of molecular interactions on the photophysics of Rose Bengal in polyelectrolyte solutions and self-assembled thin filmsMirenda, Martin; Dicelio, Lelia E.; San Roman, EnriqueJournal of Physical Chemistry B (2008), 112 (39), 12201-12207CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Aq. solns. and layer-by-layer self-assembled thin films contg. Rose Bengal and poly(diallyldimethylammonium chloride) are studied with the aim of understanding the interactions controlling their structures and the photophysics of the dye in both media. A detailed spectroscopic and theor. anal. shows that hydrophobic interactions among dye mols. contribute to the coiling of the polyelectrolyte chain in soln. at low polyelectrolyte/dye (P/D) ratios, whereas extensive aggregation of the dye takes place even at ratios as high as 104 (expressed in monomeric units). A polyelectrolyte elongated form prevails in self-assembled thin films, providing an environment that reduces hydrophobic interactions and lowers the aggregation tendency. Self-assembled films with a roughly estd. overall dye concn. around 1 M at a P/D ratio in the order of seven are fluorescent and photogenerate singlet mol. oxygen. This contrasts with the behavior of polyelectrolyte solns., which are almost nonfluorescent and do not evidence triplet state generation at the same P/D ratio.
- 29Rodríguez, H. B.; Lagorio, M. G.; San Roman, E. Rose Bengal adsorbed on microgranular cellulose: evidence on fluorescent dimers Photochem. Photobiol. Sci. 2004, 3, 674– 680 DOI: 10.1039/B402484BGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXltlynu7Y%253D&md5=36a73d53cdc4a45b8af54f12e88ad5feRose Bengal adsorbed on microgranular cellulose: evidence on fluorescent dimersRodriguez, Hernan B.; Lagorio, M. Gabriela; San Roman, EnriquePhotochemical & Photobiological Sciences (2004), 3 (7), 674-680CODEN: PPSHCB; ISSN:1474-905X. (Royal Society of Chemistry)Rose Bengal adsorbed on microgranular cellulose was studied in the solid phase by total and diffuse reflectance and steady-state emission spectroscopy. A simple self-assocn. monomer-dimer equil. fitted reflectance data up to dye loadings of 4 × 10-7 mol (g cellulose)-1 and allowed calcn. of monomer and dimer spectra. Further increase of dye loading resulted in the formation of higher aggregates. Obsd. emission and excitation spectra and quantum yields were cor. for reabsorption and reemission of luminescence, using a previously developed model, within the assumption that only monomers are luminescent. An apparent increase of fluorescence quantum yield with dye loading was found, which was attributed to the occurrence of dimer fluorescence. Extension of the model to two luminescent species (i.e. monomer and dimer) yielded const. fluorescence quantum yields for the monomer, ΦM = 0.120 ± 0.004, and for the dimer, ΦD = 0.070 ± 0.006. The monomer quantum yield is close to the value found for the same dye in basic ethanol. The presence of fluorescent dimers and calcd. quantum yields are supported by anal. of the excitation spectra and other exptl. evidence. The possible occurrence of non-radiative energy transfer and the effect of surface charge on the properties of the dimer are analyzed.
- 30Lissi, E. A.; Encinas, M. V.; Lemp, E.; Rubio, M. A. Singlet oxygen O2(1.DELTA.g) bimolecular processes. Solvent and compartmentalization effects Chem. Rev. 1993, 93, 699– 723 DOI: 10.1021/cr00018a004Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXhvFSrt7Y%253D&md5=186df1c5c33fc327b00db39d8190c2efSinglet oxygen O2(1Δg) bimolecular processes. Solvent and compartmentalization effectsLissi, E. A.; Encinas, M. V.; Lemp, E.; Rubio, M. A.Chemical Reviews (Washington, DC, United States) (1993), 93 (2), 699-723CODEN: CHREAY; ISSN:0009-2665.A review with >401 refs. including (1) deactivation dominated by energy transfer, (2) dyes and sensitizers, (3) inorg. anions, (4) N-contg. compds., (5) S-contg. compds., (6) unsatd. compds., and (7) microheterogeneous systems.
- 31Alarcón, E.; Edwards, A. M.; Aspee, A.; Moran, F. E.; Borsarelli, C. D.; Lissi, E. A.; Gonzalez-Nilo, D.; Poblete, H.; Scaiano, J. C. Photophysics and photochemistry of dyes bound to human serum albumin are determined by the dye localization Photochem. Photobiol. Sci. 2010, 9, 93– 102 DOI: 10.1039/B9PP00091GGoogle Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXht1ehsA%253D%253D&md5=ef562f3d70b318e169d5d987e421859fPhotophysics and photochemistry of dyes bound to human serum albumin are determined by the dye localizationAlarcon, Emilio; Edwards, Ana Maria; Aspee, Alexis; Moran, Faustino E.; Borsarelli, Claudio D.; Lissi, Eduardo A.; Gonzalez-Nilo, Danilo; Poblete, Horacio; Scaiano, J. C.Photochemical & Photobiological Sciences (2010), 9 (1), 93-102CODEN: PPSHCB; ISSN:1474-905X. (Royal Society of Chemistry)The photophysics and photochem. of rose bengal (RB) and methylene blue (MB) bound to human serum albumin (HSA) have been investigated under a variety of exptl. conditions. Distribution of the dyes between the external solvent and the protein has been estd. by phys. sepn. and fluorescence measurements. The main localization of protein-bound dye mols. was estd. by the intrinsic fluorescence quenching, displacement of fluorescent probes bound to specific protein sites, and by docking modeling. All the data indicate that, at low occupation nos., RB binds strongly to the HSA site I, while MB localizes predominantly in the protein binding site II. This different localization explains the obsd. differences in the dyes' photochem. behavior. In particular, the environment provided by site I is less polar and considerably less accessible to oxygen. The localization of RB in site I also leads to an efficient quenching of the intrinsic protein fluorescence (ascribed to the nearby Trp residue) and the generation of intra-protein singlet oxygen, whose behavior is different to that obsd. in the external solvent or when it is generated by bound MB.
- 32Mills, A.; Lawrence, C.; Douglas, P. Photoreduction of Water sensitised by Rose Bengal J. Chem. Soc., Faraday Trans. 2 1986, 82, 2291– 2303 DOI: 10.1039/f29868202291Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXisFKhtg%253D%253D&md5=61fb2cba0514241a10dc63b81516dc44Photoreduction of water sensitized by Rose BengalMills, Andrew; Lawrence, Carl; Douglas, PeterJournal of the Chemical Society, Faraday Transactions 2: Molecular and Chemical Physics (1986), 82 (12), 2291-303CODEN: JCFTBS; ISSN:0300-9238.Rose Bengal was used to sensitize the photoredn. of H2O in aq. EtOH soln. (5% EtOH vol./vol.; ph 4.4) via both a reductive and an oxidative quenching mechanism, using EDTA and methylviologen, resp., as quenchers of its triplet state. In the presence of EDTA (which acts as both triplet quencher and sacrificial electron donor) the photoredn. of H2O sensitized by Rose Bengal was inefficient for a no. of reasons, including a low triplet quenching rate const. (8 × 103 dm3 mol-1 s-1) and a rapid disproportionation reaction involving the protonated semi-reduced dye radicals. In the presence of methylviologen (triplet quencher) and EDTA (sacrificial electron donor) the photoredn. of H2O sensitized by Rose Bengal was inefficient due to a low cage-escape yield (.vphi. ≈ 5 × 10-2) following the primary electron-transfer step, a fast back reaction (k ≥ 1 × 108 dm3 mol-1 s-1) and formation of a complex between the dye and methylviologen (K = 6400 ± 300 cm3 mol-). The complexed form of Rose Bengal appeared able to photosensitize, inefficiently, the photoredn. of methylviologen in the presence of EDTA, without exhibiting problems of dye fade.
- 33Pupkaite, J.; Ahumada, M.; McLaughlin, S.; Temkit, M.; Alaziz, S.; Seymour, R.; Ruel, M.; Kochevar, I.; Griffith, M.; Suuronen, E. J.; Alarcon, E. I. Collagen-Based Photoactive Agent for Tissue Bonding ACS Appl. Mater. Interfaces 2017, 9, 9265– 9270 DOI: 10.1021/acsami.7b01984Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXktVSitrg%253D&md5=b64c02792565822c5f6ac13db97e9755Collagen-Based Photoactive Agent for Tissue BondingPupkaite, Justina; Ahumada, Manuel; Mclaughlin, Sarah; Temkit, Maha; Alaziz, Sura; Seymour, Richard; Ruel, Marc; Kochevar, Irene; Griffith, May; Suuronen, Erik J.; Alarcon, Emilio I.ACS Applied Materials & Interfaces (2017), 9 (11), 9265-9270CODEN: AAMICK; ISSN:1944-8244. (American Chemical Society)Using a combination of methacrylated collagen and the photosensitizer rose Bengal, a new light-activated biomimetic material for tissue sutureless bonding was developed. This formulation was crosslinked using green light. In vivo tests in mice demonstrate the suitability of the material for sutureless wound closure.
- 34Henderson, B. W.; Busch, T. M.; Vaughan, L. A.; Frawley, N. P.; Babich, D.; Sosa, T. A.; Zollo, J. D.; Dee, A. S.; Cooper, M. T.; Bellnier, D. A.; Greco, W. R.; Oseroff, A. R. Photofrin photodynamic therapy can significantly deplete or preserve oxygenation in human basal cell carcinomas during treatment, depending on fluence rate Cancer Res. 2000, 60, 525– 529Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXht1Gmtrc%253D&md5=54600117edbfbf5e2b53454a37acc235Photofrin photodynamic therapy can significantly deplete or preserve oxygenation in human basal cell carcinomas during treatment, depending on fluence rateHenderson, Barbara W.; Busch, Theresa M.; Vaughan, Lurine A.; Frawley, Noreen P.; Babich, Debra; Sosa, Tara A.; Zollo, Joseph D.; Dee, Anthony S.; Cooper, Michele T.; Bellnier, David A.; Greco, William R.; Oseroff, Allan R.Cancer Research (2000), 60 (3), 525-529CODEN: CNREA8; ISSN:0008-5472. (AACR Subscription Office)At high fluence rates in animal models, photodynamic therapy (PDT) can photochem. deplete ambient tumor oxygen through the generation of singlet oxygen, causing acute hypoxia and limiting treatment effectiveness. We report that std. clin. treatment conditions (1 mg/kg Photofrin, light at 630 nm and 150 mW/cm2), which are highly effective for treating human basal cell carcinomas, significantly diminished tumor oxygen levels during initial light delivery in a majority of carcinomas. Oxygen depletion could be found during at least 40% of the total light dose, but tumors appeared well oxygenated toward the end of treatment. In contrast, initial light delivery at a lower fluence rate of 30 mW/cm2 increased tumor oxygenation in a majority of carcinomas. Laser treatment caused an intensity- and treatment time-dependent increase in tumor temp. The data suggest that high fluence rate treatment, although effective, may be inefficient.
- 35Ravichandran, R.; Islam, M. M.; Alarcon, E. I.; Samanta, A.; Wang, S.; Lundstrom, P.; Hilborn, J.; Griffith, M.; Phopase, J. Functionalised type-I collagen as a hydrogel building block for bio-orthogonal tissue engineering applications J. Mater. Chem. B. 2016, 4, 318– 326 DOI: 10.1039/C5TB02035BGoogle Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvVyqsb3M&md5=b58d7b77544ce08f78d4e548e7b32e6bFunctionalised type-I collagen as a hydrogel building block for bio-orthogonal tissue engineering applicationsRavichandran, R.; Islam, M. M.; Alarcon, E. I.; Samanta, A.; Wang, S.; Lundstroem, P.; Hilborn, J.; Griffith, M.; Phopase, J.Journal of Materials Chemistry B: Materials for Biology and Medicine (2016), 4 (2), 318-326CODEN: JMCBDV; ISSN:2050-7518. (Royal Society of Chemistry)In this study, we derivatized type I collagen without altering its triple helical conformation to allow for facile hydrogel formation via the Michael addn. of thiols to methacrylates without the addn. of other crosslinking agents. This method provides the flexibility needed for the fabrication of injectable hydrogels or pre-fabricated implantable scaffolds, using the same components by tuning the modulus from Pa to kPa. Enzymic degradability of the hydrogels can also be easily fine-tuned by variation of the ratio and the type of the crosslinking component. The structural morphol. reveals a lamellar structure mimicking native collagen fibrils. The versatility of this material is demonstrated by its use as a pre-fabricated substrate for culturing human corneal epithelial cells and as an injectable hydrogel for 3-D encapsulation of cardiac progenitor cells.
- 36Biasini, M.; Bienert, S.; Waterhouse, A.; Arnold, K.; Studer, G.; Schmidt, T.; Kiefer, F.; Gallo Cassarino, T.; Bertoni, M.; Bordoli, L.; Schwede, T. SWISS-MODEL: modelling protein tertiary and quaternary structure using evolutionary information Nucleic Acids Res. 2014, 42, W252– W258 DOI: 10.1093/nar/gku340Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFCqs73I&md5=51509ba8353b06f286d954ebe7e6673aSWISS-MODEL: modelling protein tertiary and quaternary structure using evolutionary informationBiasini, Marco; Bienert, Stefan; Waterhouse, Andrew; Arnold, Konstantin; Studer, Gabriel; Schmidt, Tobias; Kiefer, Florian; Cassarino, Tiziano Gallo; Bertoni, Martino; Bordoli, Lorenza; Schwede, TorstenNucleic Acids Research (2014), 42 (W1), W252-W258CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)Protein structure homol. modeling has become a routine technique to generate 3D models for proteins when exptl. structures are not available. Fully automated servers such as SWISS-MODEL with user-friendly web interfaces generate reliable models without the need for complex software packages or downloading large databases. Here, we describe the latest version of the SWISS-MODEL expert system for protein structure modeling. The SWISS-MODEL template library provides annotation of quaternary structure and essential ligands and co-factors to allow for building of complete structural models, including their oligomeric structure. The improved SWISS-MODEL pipeline makes extensive use of model quality estn. for selection of the most suitable templates and provides ests. of the expected accuracy of the resulting models. The accuracy of the models generated by SWISS-MODEL is continuously evaluated by the CAMEO system. The new web site allows users to interactively search for templates, cluster them by sequence similarity, structurally compare alternative templates and select the ones to be used for model building. In cases where multiple alternative template structures are available for a protein of interest, a user-guided template selection step allows building models in different functional states. SWISS-MODEL is available at http://swissmodel.expasy.org/.
- 37Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields J. Comput. Chem. 2010, 31, 671– 690 DOI: 10.1002/jcc.21367Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtlentbc%253D&md5=26212e0e4f73bded0c89d4b411cd3833CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fieldsVanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr.Journal of Computational Chemistry (2010), 31 (4), 671-690CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The widely used CHARMM additive all-atom force field includes parameters for proteins, nucleic acids, lipids, and carbohydrates. In the present article, an extension of the CHARMM force field to drug-like mols. is presented. The resulting CHARMM General Force Field (CGenFF) covers a wide range of chem. groups present in biomols. and drug-like mols., including a large no. of heterocyclic scaffolds. The parametrization philosophy behind the force field focuses on quality at the expense of transferability, with the implementation concg. on an extensible force field. Statistics related to the quality of the parametrization with a focus on exptl. validation are presented. Addnl., the parametrization procedure, described fully in the present article in the context of the model systems, pyrrolidine, and 3-phenoxymethyl-pyrrolidine will allow users to readily extend the force field to chem. groups that are not explicitly covered in the force field as well as add functional groups to and link together mols. already available in the force field. CGenFF thus makes it possible to perform "all-CHARMM" simulations on drug-target interactions thereby extending the utility of CHARMM force fields to medicinally relevant systems. © 2009 Wiley Periodicals, Inc.J Comput Chem, 2010.
- 38Vanommeslaeghe, K.; MacKerell, A. D., Jr. Automation of the CHARMM General Force Field (CGenFF) I: bond perception and atom typing J. Chem. Inf. Model. 2012, 52, 3144– 3154 DOI: 10.1021/ci300363cGoogle Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fL&md5=c6679293f4a2501f2bcadf2020ca1473Automation of the CHARMM General Force Field (CGenFF) I: Bond Perception and Atom TypingVanommeslaeghe, K.; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3144-3154CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug-like mols. alone or interacting with biol. systems. In simulations involving biol. macromols., the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and charges is required. In the present article, which is part I of a series of two, we present the algorithms for bond perception and atom typing for the CHARMM General Force Field (CGenFF). The CGenFF atom typer first assocs. attributes to the atoms and bonds in a mol., such as valence, bond order, and ring membership among others. Of note are a no. of features that are specifically required for CGenFF. This information is then used by the atom typing routine to assign CGenFF atom types based on a programmable decision tree. This allows for straight-forward implementation of CGenFF's complicated atom typing rules and for equally straight-forward updating of the atom typing scheme as the force field grows. The presented atom typer was validated by assigning correct atom types on 477 model compds. including in the training set as well as 126 test-set mols. that were constructed to specifically verify its different components. The program may be utilized via an online implementation at https://www.paramchem.org/.
- 39Vanommeslaeghe, K.; Raman, E. P.; MacKerell, A. D., Jr. Automation of the CHARMM General Force Field (CGenFF) II: assignment of bonded parameters and partial atomic charges J. Chem. Inf. Model. 2012, 52, 3155– 3168 DOI: 10.1021/ci3003649Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fF&md5=e676ad1f42cb1e98dd353d4d285e8d13Automation of the CHARMM General Force Field (CGenFF) II: Assignment of Bonded Parameters and Partial Atomic ChargesVanommeslaeghe, K.; Raman, E. Prabhu; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3155-3168CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug candidates interacting with biol. systems. In these simulations, the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and partial at. charges is required. In the present article, algorithms for the assignment of parameters and charges for the CHARMM General Force Field (CGenFF) are presented. These algorithms rely on the existing parameters and charges that were detd. as part of the parametrization of the force field. Bonded parameters are assigned based on the similarity between the atom types that define said parameters, while charges are detd. using an extended bond-charge increment scheme. Charge increments were optimized to reproduce the charges on model compds. that were part of the parametrization of the force field. Case studies are presented to clarify the functioning of the algorithms and the significance of their output data.
- 40Best, R. B.; Zhu, X.; Shim, J.; Lopes, P. E.; Mittal, J.; Feig, M.; Mackerell, A. D., Jr. Optimization of the additive CHARMM all-atom protein force field targeting improved sampling of the backbone phi, psi and side-chain chi(1) and chi(2) dihedral angles J. Chem. Theory. Comput. 2012, 8, 3257– 3273 DOI: 10.1021/ct300400xGoogle Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtVKqurfP&md5=9a48a0c5770fb1e887c3bb34d45b1354Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone .vphi., ψ and Side-Chain χ1 and χ2 Dihedral AnglesBest, Robert B.; Zhu, Xiao; Shim, Jihyun; Lopes, Pedro E. M.; Mittal, Jeetain; Feig, Michael; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2012), 8 (9), 3257-3273CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)While the quality of the current CHARMM22/CMAP additive force field for proteins has been demonstrated in a large no. of applications, limitations in the model with respect to the equil. between the sampling of helical and extended conformations in folding simulations have been noted. To overcome this, as well as make other improvements in the model, we present a combination of refinements that should result in enhanced accuracy in simulations of proteins. The common (non-Gly, -Pro) backbone CMAP potential has been refined against exptl. soln. NMR data for weakly structured peptides, resulting in a rebalancing of the energies of the α-helix and extended regions of the Ramachandran map, correcting the α-helical bias of CHARMM22/CMAP. The Gly and Pro CMAPs have been refitted to more accurate quantum-mech. energy surfaces. Side-chain torsion parameters have been optimized by fitting to backbone-dependent quantum-mech. energy surfaces, followed by addnl. empirical optimization targeting NMR scalar couplings for unfolded proteins. A comprehensive validation of the revised force field was then performed against a collection of exptl. data: (i) comparison of simulations of eight proteins in their crystal environments with crystal structures; (ii) comparison with backbone scalar couplings for weakly structured peptides; (iii) comparison with NMR residual dipolar couplings and scalar couplings for both backbone and side-chains in folded proteins; (iv) equil. folding of mini-proteins. The results indicate that the revised CHARMM 36 parameters represent an improved model for modeling and simulation studies of proteins, including studies of protein folding, assembly, and functionally relevant conformational changes.
- 41Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kale, L.; Schulten, K. Scalable molecular dynamics with NAMD J. Comput. Chem. 2005, 26, 1781– 1802 DOI: 10.1002/jcc.20289Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SlsbbJ&md5=189051128443b547f4300a1b8fb0e034Scalable molecular dynamics with NAMDPhillips, James C.; Braun, Rosemary; Wang, Wei; Gumbart, James; Tajkhorshid, Emad; Villa, Elizabeth; Chipot, Christophe; Skeel, Robert D.; Kale, Laxmikant; Schulten, KlausJournal of Computational Chemistry (2005), 26 (16), 1781-1802CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)NAMD is a parallel mol. dynamics code designed for high-performance simulation of large biomol. systems. NAMD scales to hundreds of processors on high-end parallel platforms, as well as tens of processors on low-cost commodity clusters, and also runs on individual desktop and laptop computers. NAMD works with AMBER and CHARMM potential functions, parameters, and file formats. This article, directed to novices as well as experts, first introduces concepts and methods used in the NAMD program, describing the classical mol. dynamics force field, equations of motion, and integration methods along with the efficient electrostatics evaluation algorithms employed and temp. and pressure controls used. Features for steering the simulation across barriers and for calcg. both alchem. and conformational free energy differences are presented. The motivations for and a roadmap to the internal design of NAMD, implemented in C++ and based on Charm++ parallel objects, are outlined. The factors affecting the serial and parallel performance of a simulation are discussed. Finally, typical NAMD use is illustrated with representative applications to a small, a medium, and a large biomol. system, highlighting particular features of NAMD, for example, the Tcl scripting language. The article also provides a list of the key features of NAMD and discusses the benefits of combining NAMD with the mol. graphics/sequence anal. software VMD and the grid computing/collab. software BioCoRE. NAMD is distributed free of charge with source code at www.ks.uiuc.edu.
- 42MacKerell, A. D.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E.; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M. All-atom empirical potential for molecular modeling and dynamics studies of proteins J. Phys. Chem. B 1998, 102, 3586– 3616 DOI: 10.1021/jp973084fGoogle Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 43Feller, S. E.; Zhang, Y.; Pastor, R. W.; Brooks, B. R. Constant pressure molecular dynamics simulation: The Langevin piston method J. Chem. Phys. 1995, 103, 4613– 4621 DOI: 10.1063/1.470648Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXotVentLo%253D&md5=219a4e0a48397a35fa2c62cf99bf225aConstant pressure molecular dynamics simulation: the Langevin piston methodFeller, Scott E.; Zhang, Yuhong; Pastor, Richard W.; Brooks, Bernard R.Journal of Chemical Physics (1995), 103 (11), 4613-21CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A new method for performing mol. dynamics simulations under const. pressure is presented. In the method, which is based on the extended system formalism introduced by Andersen, the deterministic equations of motion for the piston degree of freedom are replaced by a Langevin equation; a suitable choice of collision frequency then eliminates the unphys. "ringing" of the vol. assocd. with the piston mass. In this way it is similar to the "weak coupling algorithm" developed by Berendsen and co-workers to perform mol. dynamics simulation without piston mass effects. It is shown, however, that the weak coupling algorithm induces artifacts into the simulation which can be quite severe for inhomogeneous systems such as aq. biopolymers or liq./liq. interfaces.
- 44Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N·log(N) method for Ewald sums in large systems J. Chem. Phys. 1993, 98, 10089– 10092 DOI: 10.1063/1.464397Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXks1Ohsr0%253D&md5=3c9f230bd01b7b714fd096d4d2e755f6Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 45Andersen, H. C. Rattle: A “velocity” version of the shake algorithm for molecular dynamics calculations J. Comput. Phys. 1983, 52, 24– 34 DOI: 10.1016/0021-9991(83)90014-1Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXjvFOntw%253D%253D&md5=770dfdc612edc5847839ca28ea3d6501RATTLE: a "velocity" version of the SHAKE algorithm for molecular dynamics calculationsAndersen, Hans C.Journal of Computational Physics (1983), 52 (1), 24-34CODEN: JCTPAH; ISSN:0021-9991.An algorithm, called RATTLE, for integrating the equations of motion in mol. dynamics calcns. for mol. models with internal constraints is presented. RATTLE calcs. the positions and velocities at the next time from the positions and velocities at the present time step, without requiring information about the earlier history. It is based on the Verlet algorithm and retains the simplicity of using Cartesian coordinates for each of the atoms to describe the configuration of a mol. with internal constraints. RATTLE guarantees that the coordinates and velocities of the atoms in a mol. satisfy the internal constraints at each time step.
- 46Miyamoto, S.; Kollman, P. A. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models J. Comput. Phys. 1992, 13, 952– 962 DOI: 10.1002/jcc.540130805Google ScholarThere is no corresponding record for this reference.
- 47Hopkins, C. W.; Le Grand, S.; Walker, R. C.; Roitberg, A. E. Long-Time-Step Molecular Dynamics through Hydrogen Mass Repartitioning J. Chem. Theory Comput. 2015, 11, 1864– 1874 DOI: 10.1021/ct5010406Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXksFSrsL0%253D&md5=6ca75b544b63523481de7836d11a3c1bLong-Time-Step Molecular Dynamics through Hydrogen Mass RepartitioningHopkins, Chad W.; Le Grand, Scott; Walker, Ross C.; Roitberg, Adrian E.Journal of Chemical Theory and Computation (2015), 11 (4), 1864-1874CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Previous studies have shown that the method of hydrogen mass repartitioning (HMR) is a potentially useful tool for accelerating mol. dynamics (MD) simulations. By repartitioning the mass of heavy atoms into the bonded hydrogen atoms, it is possible to slow the highest-frequency motions of the macromol. under study, thus allowing the time step of the simulation to be increased by up to a factor of 2. In this communication, we investigate further how this mass repartitioning allows the simulation time step to be increased in a stable fashion without significantly increasing discretization error. To this end, we ran a set of simulations with different time steps and mass distributions on a three-residue peptide to get a comprehensive view of the effect of mass repartitioning and time step increase on a system whose accessible phase space is fully explored in a relatively short amt. of time. We next studied a 129-residue protein, hen egg white lysozyme (HEWL), to verify that the obsd. behavior extends to a larger, more-realistic, system. Results for the protein include structural comparisons from MD trajectories, as well as comparisons of pKa calcns. via const.-pH MD. We also calcd. a potential of mean force (PMF) of a dihedral rotation for the MTS [(1-oxyl-2,2,5,5-tetramethyl-pyrroline-3-methyl)methanethiosulfonate] spin label via umbrella sampling with a set of regular MD trajectories, as well as a set of mass-repartitioned trajectories with a time step of 4 fs. Since no significant difference in kinetics or thermodn. is obsd. by the use of fast HMR trajectories, further evidence is provided that long-time-step HMR MD simulations are a viable tool for accelerating MD simulations for mols. of biochem. interest.
- 48Fiorin, G.; Klein, M. L.; Hénin, J. Using collective variables to drive molecular dynamics simulations Mol. Phys. 2013, 111, 3345– 3362 DOI: 10.1080/00268976.2013.813594Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXpslGit7s%253D&md5=6c1886ae6a2804383260577562fb503bUsing collective variables to drive molecular dynamics simulationsFiorin, Giacomo; Klein, Michael L.; Henin, JeromeMolecular Physics (2013), 111 (22-23), 3345-3362CODEN: MOPHAM; ISSN:0026-8976. (Taylor & Francis Ltd.)A software framework is introduced that facilitates the application of biasing algorithms to collective variables of the type commonly employed to drive massively parallel mol. dynamics (MD) simulations. The modular framework that is presented enables one to combine existing collective variables into new ones, and combine any chosen collective variable with available biasing methods. The latter include the classic time-dependent biases referred to as steered MD and targeted MD, the temp.-accelerated MD algorithm, as well as the adaptive free-energy biases called metadynamics and adaptive biasing force. The present modular software is extensible, and portable between commonly used MD simulation engines.
- 49Comer, J.; Gumbart, J. C.; Hénin, J.; Lelièvre, T.; Pohorille, A.; Chipot, C. The Adaptive Biasing Force Method: Everything You Always Wanted To Know but Were Afraid To Ask J. Phys. Chem. B 2015, 119, 1129– 1151 DOI: 10.1021/jp506633nGoogle Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhsF2nurvP&md5=2ea7fa53559bf7355d24630fc126fcc9The Adaptive Biasing Force Method: Everything You Always Wanted To Know but Were Afraid To AskComer, Jeffrey; Gumbart, James C.; Henin, Jerome; Lelievre, Tony; Pohorille, Andrew; Chipot, ChristopheJournal of Physical Chemistry B (2015), 119 (3), 1129-1151CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)In the host of numerical schemes devised to calc. free energy differences by way of geometric transformations, the adaptive biasing force algorithm has emerged as a promising route to map complex free-energy landscapes. It relies upon the simple concept that as a simulation progresses, a continuously updated biasing force is added to the equations of motion, such that in the long-time limit it yields a Hamiltonian devoid of an av. force acting along the transition coordinate of interest. This means that sampling proceeds uniformly on a flat free-energy surface, thus providing reliable free-energy ests. Much of the appeal of the algorithm to the practitioner is in its phys. intuitive underlying ideas and the absence of any requirements for prior knowledge about free-energy landscapes. Since its inception in 2001, the adaptive biasing force scheme has been the subject of considerable attention, from in-depth math. anal. of convergence properties to novel developments and extensions. The method has also been successfully applied to many challenging problems in chem. and biol. In this contribution, the method is presented in a comprehensive, self-contained fashion, discussing with a crit. eye its properties, applicability, and inherent limitations, as well as introducing novel extensions. Through free-energy calcns. of prototypical mol. systems, many methodol. aspects are examd., from stratification strategies to overcoming the so-called hidden barriers in orthogonal space, relevant not only to the adaptive biasing force algorithm but also to other importance-sampling schemes. On the basis of the discussions in this paper, a no. of good practices for improving the efficiency and reliability of the computed free-energy differences are proposed.
- 50Darve, E.; Rodriguez-Gomez, D.; Pohorille, A. Adaptive biasing force method for scalar and vector free energy calculations J. Chem. Phys. 2008, 128144120 DOI: 10.1063/1.2829861Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXkvFyiu74%253D&md5=6f6eb47d685e873d1ff35ffdc9ae66cbAdaptive biasing force method for scalar and vector free energy calculationsDarve, Eric; Rodriguez-Gomez, David; Pohorille, AndrewJournal of Chemical Physics (2008), 128 (14), 144120/1-144120/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)In free energy calcns. based on thermodn. integration, it is necessary to compute the derivs. of the free energy as a function of one (scalar case) or several (vector case) order parameters. We derive in a compact way a general formulation for evaluating these derivs. as the av. of a mean force acting on the order parameters, which involves first derivs. with respect to both Cartesian coordinates and time. This is in contrast with the previously derived formulas, which require first and second derivs. of the order parameter with respect to Cartesian coordinates. As illustrated in a concrete example, the main advantage of this new formulation is the simplicity of its use, esp. for complicated order parameters. It is also straightforward to implement in a mol. dynamics code, as can be seen from the pseudo-code given at the end. We further discuss how the approach based on time derivs. can be combined with the adaptive biasing force method, an enhanced sampling technique that rapidly yields uniform sampling of the order parameters, and by doing so greatly improves the efficiency of free energy calcns. Using the backbone dihedral angles Φ and Ψ in N-acetylalanyl-N'-methylamide as a numerical example, we present a technique to reconstruct the free energy from its derivs., a calcn. that presents some difficulties in the vector case because of the statistical errors affecting the derivs. (c) 2008 American Institute of Physics.
- 51Hénin, J.; Chipot, C. Overcoming free energy barriers using unconstrained molecular dynamics simulations J. Chem. Phys. 2004, 121, 2904– 2914 DOI: 10.1063/1.1773132Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXmt1Smurk%253D&md5=1f5935c22099d8a213df1249c8ace7bfOvercoming free energy barriers using unconstrained molecular dynamics simulationsHenin, Jerome; Chipot, ChristopheJournal of Chemical Physics (2004), 121 (7), 2904-2914CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Assocn. of unconstrained mol. dynamics (MD) and the formalisms of thermodn. integration and av. force [Darve and Pohorille, J. Chem. Phys. 115, 9169 (2001)] have been employed to det. potentials of mean force. When implemented in a general MD code, the addnl. computational effort, compared to other std., unconstrained simulations, is marginal. The force acting along a chosen reaction coordinate ξ is estd. from the individual forces exerted on the chem. system and accumulated as the simulation progresses. The estd. free energy deriv. computed for small intervals of ξ is canceled by an adaptive bias to overcome the barriers of the free energy landscape. Evolution of the system along the reaction coordinate is, thus, limited by its sole self-diffusion properties. The illustrative examples of the reversible unfolding of deca-L-alanine, the assocn. of acetate and guanidinium ions in water, the dimerization of methane in water, and its transfer across the water liq.-vapor interface are examd. to probe the efficiency of the method.
Cited By
This article is cited by 38 publications.
- Vivian M. Lechner, Manuel Nappi, Patrick J. Deneny, Sarah Folliet, John C. K. Chu, Matthew J. Gaunt. Visible-Light-Mediated Modification and Manipulation of Biomacromolecules. Chemical Reviews 2022, 122
(2)
, 1752-1829. https://doi.org/10.1021/acs.chemrev.1c00357
- Christopher D. McTiernan, David C. Cortes, Caitlin Lazurko, Selya Amrani, Roberto Rosales-Rojas, Matias Zuñiga-Bustos, Veronika Sedlakova, Horacio Poblete, Kevin Stamplecoskie, Erik J. Suuronen, Emilio I. Alarcon. Light-Activated Peptide-Based Materials for Sutureless Wound Closure. ACS Applied Materials & Interfaces 2019, 11
(48)
, 45007-45015. https://doi.org/10.1021/acsami.9b18891
- Herleen Ruprai, Sara Romanazzo, Jake Ireland, Kristopher Kilian, Damia Mawad, Laurel George, Richard Wuhrer, Jessica Houang, Daniel Ta, Simon Myers, Antonio Lauto. Porous Chitosan Films Support Stem Cells and Facilitate Sutureless Tissue Repair. ACS Applied Materials & Interfaces 2019, 11
(36)
, 32613-32622. https://doi.org/10.1021/acsami.9b09123
- Lucie Ludvíková, Peter Štacko, Jonathan Sperry, Petr Klán. Photosensitized Cross-Linking of Tryptophan and Tyrosine Derivatives by Rose Bengal in Aqueous Solutions. The Journal of Organic Chemistry 2018, 83
(18)
, 10835-10844. https://doi.org/10.1021/acs.joc.8b01545
- Ning Chai, Tanja Stachon, Tim Berger, Zhen Li, Berthold Seitz, Achim Langenbucher, Nóra Szentmáry. Short-Term Effect of Rose Bengal Photodynamic Therapy (RB-PDT) on Collagen I, Collagen V, NF-κB, LOX, TGF-β and IL-6 Expression of Human Corneal Fibroblasts,
In Vitro. Current Eye Research 2023, 43 , 1-8. https://doi.org/10.1080/02713683.2023.2276057
- R. Haindl, V. Bellemo, P. Rajendran, B. Tan, M. Liu, B. S. Lee, Q. Zhou, R. A. Leitgeb, W. Drexler, L. Schmetterer, M. Pramanik. Visible light photoacoustic ophthalmoscopy and near-infrared-II optical coherence tomography in the mouse eye. APL Photonics 2023, 8
(10)
https://doi.org/10.1063/5.0168091
- Rongrong Gao, Ming Chen, Xueyang Chen, Xinyu Liu, Qingqing Jiang, Keith M. Meek, Qinmei Wang, Shihao Chen, Jinhai Huang. Diffusion Depth and Efficacy of Different Infiltration Times for Rose Bengal/Green Light Corneal Cross-linking in Rabbit Eyes. Journal of Refractive Surgery 2023, 39
(9)
, 620-626. https://doi.org/10.3928/1081597X-20230726-01
- Claudia Cecilia Vera, Claudio Darío Borsarelli. Photo-induced protein modifications: a range of biological consequences and applications. Biophysical Reviews 2023, 15
(4)
, 569-576. https://doi.org/10.1007/s12551-023-01081-6
- Constanza Alvarez‐Lopez, Daniela Cavazos‐Elizondo, Belinda Heyne, Irene E. Kochevar, Alan Aguirre‐Soto. Nanocaging Rose Bengal to Inhibit Aggregation and Enhance Photo‐induced Oxygen Consumption
†. Photochemistry and Photobiology 2023, 99
(2)
, 580-592. https://doi.org/10.1111/php.13746
- Ning Chai, Tanja Stachon, Mahsa Nastaranpour, Zhen Li, Berthold Seitz, Myriam Ulrich, Achim Langenbucher, Nóra Szentmáry. Assessment of Rose Bengal Photodynamic Therapy on Viability and Proliferation of Human Keratolimbal Epithelial and Stromal Cells In Vitro. Klinische Monatsblätter für Augenheilkunde 2023, 53 https://doi.org/10.1055/a-2038-8899
- Michelle Fiona Kilb, Ulrike Ritz, Daniela Nickel, Katja Schmitz. pH-Dependent Release of Vancomycin from Modularly Assembled Collagen Laminates. Polymers 2022, 14
(23)
, 5227. https://doi.org/10.3390/polym14235227
- Xin Zhou, Zhongzhu Yang, Yu Chen, Haopeng Feng, Jiangfang Yu, Jialin Tang, Xiaoyi Ren, Jing Tang, Jiajia Wang, Lin Tang. Single-atom Ru loaded on layered double hydroxide catalyzes peroxymonosulfate for effective E. coli inactivation via a non-radical pathway: Efficiency and mechanism. Journal of Hazardous Materials 2022, 440 , 129720. https://doi.org/10.1016/j.jhazmat.2022.129720
- Rongrong Gao, Mengdi Yan, Ming Chen, Sally Hayes, Keith M. Meek, Huanhuan He, Xueyang Chen, Wenjin Xu, Shixiang Yan, Yuyan Huang, Shengnan Ding, Qinmei Wang, Junhua Li, Jinhai Huang. The Impact of Different Rose Bengal Formulations on Corneal Thickness and the Efficacy of Rose Bengal/Green Light Corneal Cross-linking in the Rabbit Eye. Journal of Refractive Surgery 2022, 38
(7)
, 450-458. https://doi.org/10.3928/1081597X-20220601-03
- Emille M. Rodrigues, Nicholas D. Calvert, Justin C. Crawford, Nan Liu, Adam J. Shuhendler, Eva Hemmer. Phytoglycogen Encapsulation of Lanthanide‐Based Nanoparticles as an Optical Imaging Platform with Therapeutic Potential. Small 2022, 18
(24)
https://doi.org/10.1002/smll.202107130
- Roberto Rosales-Rojas, Matías Zuñiga-Bustos, Francisca Salas-Sepúlveda, Constanza Galaz-Araya, Ricardo A. Zamora, Horacio Poblete. Self-Organization Dynamics of Collagen-like Peptides Crosslinking Is Driven by Rose-Bengal-Mediated Electrostatic Bridges. Pharmaceutics 2022, 14
(6)
, 1148. https://doi.org/10.3390/pharmaceutics14061148
- Chengfeng Cai, Jing-Ju Huang, Koki Sano, Yicheng Zhu, Yunhua Zhang, Qing Wu, Bo Zhu, Yasuhiro Ishida, Shyh-Chyang Luo, Hsing-An Lin. A water-soluble corannulene with highly efficient ROS production. Materials Chemistry and Physics 2022, 281 , 125885. https://doi.org/10.1016/j.matchemphys.2022.125885
- Yi Liu, Wenhua Dong, Shuyi Shen, Fanxu Meng, Jiajia Wang, Kun Yang, Daohui Lin. Enhancement of E. coli inactivation by photosensitized erythrosine-based solar disinfection under weakly acidic conditions. Water Research 2022, 212 , 118125. https://doi.org/10.1016/j.watres.2022.118125
- Yanna Gurianov, Michael Meistelman, Yael Albo, Marina Nisnevitch, Faina Nakonechny. Antibacterial Activity of Rose Bengal Entrapped in Organically Modified Silica Matrices. International Journal of Molecular Sciences 2022, 23
(7)
, 3716. https://doi.org/10.3390/ijms23073716
- Constanza Alvarez-López, Alan Aguirre-Soto. A review on Light-initiated crosslinking of biomaterials stained with Rose Bengal for “Nanosuturing” Macromolecular interfaces. Materials Today: Proceedings 2022, 48 , 56-65. https://doi.org/10.1016/j.matpr.2020.10.159
- Olga Kravchenko, Todd C. Sutherland, Belinda Heyne. Photobleaching of Erythrosine B in Aqueous Environment Investigation Beyond pH
†. Photochemistry and Photobiology 2022, 98
(1)
, 49-56. https://doi.org/10.1111/php.13396
- Farnoush Sadat Rezaei, Fariborz Sharifianjazi, Amirhossein Esmaeilkhanian, Ehsan Salehi. Chitosan films and scaffolds for regenerative medicine applications: A review. Carbohydrate Polymers 2021, 273 , 118631. https://doi.org/10.1016/j.carbpol.2021.118631
- Eugenio Hernan Otal, Manuela Leticia Kim, Yoshiyuki Hattori, Yu Kitazawa, Juan Paulo Hinestroza, Mutsumi Kimura. Versatile Covalent Postsynthetic Modification of Metal Organic Frameworks via Thermal Condensation for Fluoride Sensing in Waters. Bioengineering 2021, 8
(12)
, 196. https://doi.org/10.3390/bioengineering8120196
- Hakan Ceylan, Nihal Olcay Dogan, Immihan Ceren Yasa, Mirac Nur Musaoglu, Zeynep Umut Kulali, Metin Sitti. 3D printed personalized magnetic micromachines from patient blood–derived biomaterials. Science Advances 2021, 7
(36)
https://doi.org/10.1126/sciadv.abh0273
- Jaime D. Martinez, Esdras Arrieta, Andrea Naranjo, Pedro Monsalve, Keenan J. Mintz, Jeffrey Peterson, Alejandro Arboleda, Heather Durkee, Mariela C. Aguilar, Daniel Pelaez, Sander R. Dubovy, Darlene Miller, Roger Leblanc, Guillermo Amescua, Jean-Marie Parel. Rose Bengal Photodynamic Antimicrobial Therapy: A Pilot Safety Study. Cornea 2021, 40
(8)
, 1036-1043. https://doi.org/10.1097/ICO.0000000000002717
- Bryan Mendes, Stefan Kassumeh, Alan Aguirre‐Soto, Qing Pei, Belinda Heyne, Irene E. Kochevar. Influence of Rose Bengal Dimerization on Photosensitization. Photochemistry and Photobiology 2021, 97
(4)
, 718-726. https://doi.org/10.1111/php.13379
- Keun Ah Ryu, Christina M. Kaszuba, Noah B. Bissonnette, Rob C. Oslund, Olugbeminiyi O. Fadeyi. Interrogating biological systems using visible-light-powered catalysis. Nature Reviews Chemistry 2021, 5
(5)
, 322-337. https://doi.org/10.1038/s41570-021-00265-6
- Soohyun Jeong, Sujin Kim, Youngjoo Choi, Han Na Jung, Kangwon Lee, Min Hee Park, . Development of Glycerol-Rose Bengal-Polidocanol (GRP) foam for enhanced sclerosis of a cyst for cystic diseases. PLOS ONE 2021, 16
(1)
, e0244635. https://doi.org/10.1371/journal.pone.0244635
- Jeffrey C. Peterson, Esdras Arrieta, Marco Ruggeri, Juan D. Silgado, Keenan J. Mintz, Ernesto H. Weisson, Roger M. Leblanc, Irene Kochevar, Fabrice Manns, Jean-Marie Parel. Detection of singlet oxygen luminescence for experimental corneal rose bengal photodynamic antimicrobial therapy. Biomedical Optics Express 2021, 12
(1)
, 272. https://doi.org/10.1364/BOE.405601
- Christian M. Wertheimer, Bryan Mendes, Qing Pei, Katharina Brandt, Irene E. Kochevar. Arginine as an Enhancer in Rose Bengal Photosensitized Corneal Crosslinking. Translational Vision Science & Technology 2020, 9
(8)
, 24. https://doi.org/10.1167/tvst.9.8.24
- Nicholas Gillman, David Lloyd, Randy Bindra, Rui Ruan, Minghao Zheng. Surgical applications of intracorporal tissue adhesive agents: current evidence and future development. Expert Review of Medical Devices 2020, 17
(5)
, 443-460. https://doi.org/10.1080/17434440.2020.1743682
- Eduardo Fuentes-Lemus, Camilo López-Alarcón. Photo-induced protein oxidation: mechanisms, consequences and medical applications. Essays in Biochemistry 2020, 64
(1)
, 33-44. https://doi.org/10.1042/EBC20190044
- Herleen Ruprai, Anu Shanu, Damia Mawad, James M. Hook, Kristopher Kilian, Laurel George, Richard Wuhrer, Jessica Houang, Simon Myers, Antonio Lauto. Porous chitosan adhesives with L-DOPA for enhanced photochemical tissue bonding. Acta Biomaterialia 2020, 101 , 314-326. https://doi.org/10.1016/j.actbio.2019.10.046
- Eduardo Fuentes-Lemus, Michele Mariotti, Per Hägglund, Fabian Leinisch, Angélica Fierro, Eduardo Silva, Camilo López-Alarcón, Michael J. Davies. Binding of rose bengal to lysozyme modulates photooxidation and cross-linking reactions involving tyrosine and tryptophan. Free Radical Biology and Medicine 2019, 143 , 375-386. https://doi.org/10.1016/j.freeradbiomed.2019.08.023
- Stacey Sova, Lisa A. Kelly. Enzyme Modification and Oxidative Cross‐linking Using Carboxylate‐, Phenol‐ and Catechol‐Conjugated 1,8‐Naphthalimides. Photochemistry and Photobiology 2019, 95
(5)
, 1169-1178. https://doi.org/10.1111/php.13110
- Robert W. Redmond, Irene E. Kochevar. Medical Applications of Rose Bengal‐ and Riboflavin‐Photosensitized Protein Crosslinking. Photochemistry and Photobiology 2019, 95
(5)
, 1097-1115. https://doi.org/10.1111/php.13126
- Christian M. Wertheimer, Carolin Elhardt, Steffen M. Kaminsky, Linh Pham, Qing Pei, Bryan Mendes, Sepideh Afshar, Irene E. Kochevar. Enhancing Rose Bengal-Photosensitized Protein Crosslinking in the Cornea. Investigative Opthalmology & Visual Science 2019, 60
(6)
, 1845. https://doi.org/10.1167/iovs.19-26604
- Robert D. Bongard, Michael Lepley, Adam Gastonguay, Raulia R. Syrlybaeva, Marat R. Talipov, Rachel A. Jones Lipinski, Noah R. Leigh, Jaladhi Brahmbhatt, Raman Kutty, Rajendra Rathore, Ramani Ramchandran, Daniel S. Sem. Discovery and characterization of halogenated xanthene inhibitors of DUSP5 as potential photodynamic therapeutics. Journal of Photochemistry and Photobiology A: Chemistry 2019, 375 , 114-131. https://doi.org/10.1016/j.jphotochem.2019.01.005
- Alan Aguirre-Soto, Seunghyeon Kim, Kaja Kaastrup, Hadley D. Sikes. On the role of
N
-vinylpyrrolidone in the aqueous radical-initiated copolymerization with PEGDA mediated by eosin Y in the presence of O
2. Polymer Chemistry 2019, 10
(8)
, 926-937. https://doi.org/10.1039/C8PY01459K
Abstract
Figure 1
Figure 1. Effect of collagen-like peptide on 10 μM rose Bengal absorption. (a) Spectral changes and (b) changes in the absorption at RB monomer maximum 550 nm and at 570 nm, the maximum of the red-shifted peak for aggregated RB bound to CLP, plotted as a function of CLP concentration. All measurements were carried out in 10 mM MES buffer (pH 5.0).
Figure 2
Figure 2. Right: Normalized absorption spectra for RB prepared at different dye concentrations incorporated within collagen matrices. Left: Normalized absorption spectra for RB measured in buffer solutions. All measurements were carried out in phosphate buffer pH 7.4 at room temperature.
Figure 3
Figure 3. Molecular simulation of RB binding to the CLP. (a) Molecular model of RB in the charge state used in the simulations. The carboxylate and phenolate moieties each have charges of −1. Atom color code: H, white; C, cyan; O, red; Cl, green; I, purple. (b) Snapshot of a simulation of the CLP triple helix with one molecule of RB. The CLP is shown as a green tube, Na+ and Cl– ions are shown as yellow and cyan spheres. For clarity, the explicit water molecules are shown as a transparent surface. (c) Free-energy landscape for RB in the vicinity of the CLP triple helix at low RB concentration, calculated by the adaptive biasing force method. The potential of mean force is mapped as a function of the position of the RB molecule along the CLP axis (disZ) and distance from this axis (disXY). The geometric contribution to the free energy along disXY has been removed. (d) Free-energy landscape at a higher RB concentration calculated from a set of equilibrium simulations including a CLP triple helix and 20 RB molecules. (e) Snapshot from the simulation with 20 RB molecules.
Figure 4
Figure 4. Molecular interactions between RB and CLP. (a) Electrostatic contact and H-bonding between the N-terminus of the CLP and the carboxylate group of RB. In these images, the phenolate (O–) group is highlighted by a red sphere to distinguish it from the carbonyl oxygen. Atoms are colored as in Figure 3, except that CLP carbons are colored dark green. H-bonds are indicated by dotted black lines. (b) Hydrophobic contact between the three-ring moiety of RB and a Pro residue of the CLP. (c) H-bonding between the O– group of RB and the backbone NH of the CLP, coupled with ionic contact between the O– group and a Lys side chain. (d) H-bonding between the OH group of hydroxyproline (Hyp) and the carboxylate group of RB. An ionic Lys–O– contact is also apparent. (e) The free-energy landscape of RB near the CLP with labels (a–d) indicating the location of RB in the corresponding panels of this figure. The ranges of disZ values occupied by the N-terminus and Asp, Hyp, and Lys side chains are also indicated. (f) The prevalence of H-bonding between RB and different groups of the CLP as a function of position along the CLP axis. Plotted are the number of H-bonds involving all CLP groups (CLP), backbone amide nitrogens (BB), N-terminal NH3+ groups (Nter), Hyp side chain OH groups, and Lys side chain NH3+ groups. (g) The prevalence of H-bonding between the CLP and different groups of RB as a function of position along the CLP axis.
Figure 5
Figure 5. Photodecomposition of rose Bengal (125 μM) in the presence of arginine (25 mM). (a) Irradiation of rose Bengal in nitrogen saturated solution. (b) Absorption spectra (expanded scale) of rose Bengal samples irradiated for 7 min in the presence of 25 mM Arg in solutions saturated with oxygen, air, and nitrogen. All data collected in PBS buffer pH 7.4 at room temperature. Solutions irradiated at 532 nm, irradiance = 0.3 W/cm2. Results shown are from single irradiations in triplicate independent experiments.
Figure 6
Figure 6. Photodegradation of 10 μM RB in the presence of different additives: (a) without and (b) with 2.5 μM collagen at RB/collagen (4:1). Left: Absorption spectra for RB measured in solution at different irradiation times. Right: Changes in absorption intensities measured at 550 nm, plotted as A/A0, for RB solutions recorded in the presence of different additives. All data collected in 10 mM MES buffer pH 5.0 at room temperature. Data correspond to the average calculated from three independent experiments each repeated in triplicate (n = 9).
Figure 7
Figure 7. Photodegradation of 40 μM RB incorporated within collagen hydrogels prepared using type I collagen containing different additives. Left: Absorption spectra for the RB–collagen hydrogel composite measured at different irradiation times in an air atmosphere. Right: Changes in absorption intensities measured at 556 nm for RB solutions recorded in the presence of different additives [Azide: 10 mM, Trp: 2 mM, Arg: 10 mM]. All data collected in PBS buffer pH 7.4 at room temperature. Data correspond to the average calculated from three independent experiments each repeated in triplicate (n = 9).
Figure 8
Figure 8. Triplet transient lifetime and quenching of RB embedded in collagen hydrogels in the presence of different quenchers. (a) RB triplet decay monitored at 620 nm measured in air (black circles) or nitrogen (blue circles) saturated solutions. Insets correspond to decay residuals obtained from the exponential fit for the decays shown in the figure. Measured lifetimes for: (b) RB triplet, (c) RB anion radical, and (d) singlet oxygen under different conditions or additives. Concentrations of additives were: tryptophan: 2.0 mM, Arg: 10 mM, and sodium azide: 10 mM. Effect of additives for laser flash photolysis experiments were carried out in nitrogen saturated solutions. For singlet oxygen measurements, samples were equilibrated with air prior to laser excitation. Measurements for triplet and anion radical were carried out in 10 mM pH 5.0 MES buffer. For singlet oxygen phosphorescence, a 10 mM MES buffer was prepared with a pD of 5.0. In all cases, photodegradation was kept lower than 10%. Time traces correspond to the average of 12 separate decays from 3 independent samples.
References
ARTICLE SECTIONSThis article references 51 other publications.
- 1Chan, B. P.; Amann, C.; Yaroslavsky, A. N.; Title, C.; Smink, D.; Zarins, B.; Kochevar, I. E.; Redmond, R. W. Photochemical repair of Achilles tendon rupture in a rat model J. Surg. Res. 2005, 124, 274– 279 DOI: 10.1016/j.jss.2004.09.019Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXjtVyrsrc%253D&md5=7021988a830cf8f0b6a4a21ef0a0f492Photochemical repair of Achilles tendon rupture in a rat modelChan, Barbara P.; Amann, Christopher; Yaroslavsky, Anna N.; Title, Craig; Smink, David; Zarins, Bertram; Kochevar, Irene E.; Redmond, Robert W.Journal of Surgical Research (2005), 124 (2), 274-279CODEN: JSGRA2; ISSN:0022-4804. (Elsevier)Photochem. tissue bonding (PTB) is an emerging technique for bonding or sealing tissue surfaces that requires light and a photoactive dye for its effect. The potential of PTB for tendon repair was assessed in a rat model. The optical properties of bovine tendon were detd. ex vivo to gauge the depth of light penetration as a function of wavelength and dosimetry parameters were established for PTB repair of ruptured tendon. PTB was then tested in vivo to repair transected tendons in Sprague-Dawley rats. Repair strengths were measured using a strain gauge up to 14 days post treatment. The effective penetration depth in tendon was estd. to be 0.68 mm at 514 nm. Following PTB treatment of mech. ruptured tendon, significant bonding was dependent on the presence of both light and dye and attained a plateau strength at a fluence of 125 J/cm2. In a subsequent in vivo study to investigate PTB for repair of transected rat Achilles tendon, the ultimate stress required to break the repaired tendon was measured immediately after irradn. and at 7 and 14 days post-repair. Results showed that the difference in the ultimate stress between control and PTB treatment groups was statistically significant immediately after treatment and at 7 days (p = 0.04) but not 14 days (p = 0.75) post-repair. PTB provides a benefit to tendon repair at early stages in repair and is worthy of further investigation as a potential surgical adjunct for tendon repair in orthopedic surgeries.
- 2Fairbairn, N. G.; Ng-Glazier, J.; Meppelink, A. M.; Randolph, M. A.; Valerio, I. L.; Fleming, M. E.; Kochevar, I. E.; Winograd, J. M.; Redmond, R. W. Light-Activated Sealing of Acellular Nerve Allografts following Nerve Gap Injury J. Reconstr. Microsurg. 2016, 32, 421– 430 DOI: 10.1055/s-0035-1571247Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28jgsleqtw%253D%253D&md5=0c21b1b08b0e14cfd2d3fb2fd3ad5ffcLight-Activated Sealing of Acellular Nerve Allografts following Nerve Gap InjuryFairbairn Neil G; Ng-Glazier Joanna; Meppelink Amanda M; Randolph Mark A; Winograd Jonathan M; Valerio Ian L; Fleming Mark E; Kochevar Irene E; Redmond Robert WJournal of reconstructive microsurgery (2016), 32 (6), 421-30 ISSN:.Introduction Photochemical tissue bonding (PTB) uses visible light to create sutureless, watertight bonds between two apposed tissue surfaces stained with photoactive dye. In phase 1 of this two-phase study, nerve gaps repaired with bonded isografts were superior to sutured isografts. When autograft demand exceeds supply, acellular nerve allograft (ANA) is an alternative although outcomes are typically inferior. This study assesses the efficacy of PTB when used with ANA. Methods Overall 20 male Lewis rats had 15-mm left sciatic nerve gaps repaired using ANA. ANAs were secured using epineurial suture (group 1) or PTB (group 2). Outcomes were assessed using sciatic function index (SFI), gastrocnemius muscle mass retention, and nerve histomorphometry. Historical controls from phase 1 were used to compare the performance of ANA with isograft. Statistical analysis was performed using analysis of variance and Bonferroni all-pairs comparison. Results All ANAs had signs of successful regeneration. Mean values for SFI, muscle mass retention, nerve fiber diameter, axon diameter, and myelin thickness were not significantly different between ANA + suture and ANA + PTB. On comparative analysis, ANA + suture performed significantly worse than isograft + suture from phase 1. However, ANA + PTB was statistically comparable to isograft + suture, the current standard of care. Conclusion Previously reported advantages of PTB versus suture appear to be reduced when applied to ANA. The lack of Schwann cells and neurotrophic factors may be responsible. PTB may improve ANA performance to an extent, where they are equivalent to autograft. This may have important clinical implications when injuries preclude the use of autograft.
- 3Mulroy, L.; Kim, J.; Wu, I.; Scharper, P.; Melki, S. A.; Azar, D. T.; Redmond, R. W.; Kochevar, I. E. Photochemical keratodesmos for repair of lamellar corneal incisions Invest. Ophthalmol. Visual Sci. 2000, 41, 3335– 3340Google ScholarThere is no corresponding record for this reference.
- 4O’Neill, A. C.; Winograd, J. M.; Zeballos, J. L.; Johnson, T. S.; Randolph, M. A.; Bujold, K. E.; Kochevar, I. E.; Redmond, R. W. Microvascular anastomosis using a photochemical tissue bonding technique Lasers Surg. Med. 2007, 39, 716– 722 DOI: 10.1002/lsm.20548Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BD2snkvVehsQ%253D%253D&md5=0d0b5e9d0759a9e621613c966539ded3Microvascular anastomosis using a photochemical tissue bonding techniqueO'Neill Anne C; Winograd Jonathan M; Zeballos Jose L; Johnson T Shane; Randolph Mark A; Bujold Kenneth E; Kochevar Irene E; Redmond Robert WLasers in surgery and medicine (2007), 39 (9), 716-22 ISSN:0196-8092.BACKGROUND AND OBJECTIVES: Photochemical tissue bonding (PTB) combines photoactive dyes with visible light to create fluid-tight seals between tissue surfaces without causing collateral thermal damage. The potential of PTB to improve outcomes over standard of care microsurgical reanastomoses of blood vessels in ex vivo and in vivo models was evaluated. STUDY DESIGN: The mechanical strength and integrity of PTB and standard microsurgical suture repairs in ex vivo porcine brachial arteries (n = 10) were compared using hydrostatic testing of leak point pressure (LPP). Femoral artery repair in vivo was measured in Sprague-Dawley rats using either standard microvascular sutures (n = 20) or PTB (n = 20). Patency was evaluated at 6 hours (n = 10) and 8 weeks post-repair (n = 10) for each group. RESULTS: PTB produced significantly higher LPPs (1,100+/- 150 mmHg) than suture repair (350+/-40 mmHg, P<0.001) in an ex vivo study. In an in vivo study all femoral arteries in both suture and PTB repair groups were patent at 6 hours post-repair. At 8 weeks post-repair the patency rate was 80% for both groups. No evidence of aneurysm formation was seen in either group and bleeding was absent from the repair site in the PTB-treated vessels, in contrast to the suture repair group. CONCLUSION: PTB is a feasible microvascular repair technique that results in an immediate, mechanically robust bond with short- and long-term patency rates equal to those for standard suture repair.
- 5Senthil-Kumar, P.; Ni, T.; Randolph, M. A.; Velmahos, G. C.; Kochevar, I. E.; Redmond, R. W. A light-activated amnion wrap strengthens colonic anastomosis and reduces peri-anastomotic adhesions Lasers Surg. Med. 2016, 48, 530– 537 DOI: 10.1002/lsm.22507Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28fitVWgug%253D%253D&md5=c2ef2ce184a37b1f44bfbafe4dd08bb8A light-activated amnion wrap strengthens colonic anastomosis and reduces peri-anastomotic adhesionsSenthil-Kumar Prabhu; Ni Tao; Kochevar Irene E; Redmond Robert W; Senthil-Kumar Prabhu; Randolph Mark A; Ni Tao; Velmahos George CLasers in surgery and medicine (2016), 48 (5), 530-7 ISSN:.BACKGROUND AND OBJECTIVE: Colonic anastomotic failure is a dreaded complication, and multiple surgical techniques have failed to eliminate it. Photochemical tissue bonding (PTB) is a method of sealing tissue surfaces by light-activated crosslinking. We evaluated if a human amniotic membrane (HAM), sealed over the anastomotic line by PTB, increases the anastomotic strength. STUDY DESIGN: Sprague-Dawley rats underwent midline laparotomy followed by surgical transection of the left colon. Animals were randomized to colonic anastomosis by one of the following methods (20 per group): single-layer continuous circumferential suture repair (SR); SR with a HAM wrap attached by suture (SR+ HAM-S); SR with HAM bonded photochemically over the anastomotic site using 532 nm light (SR+ HAM-PTB); approximation of the bowel ends with only three sutures and sealing with HAM-PTB (3+ HAM-PTB). A control group underwent laparotomy alone with no colon resection (NR). Sub-groups (n = 10) were sacrificed at days 3 and 7 post-operatively and adhesions were evaluated. A 6 cm section of colon was then removed and strength of anastomosis evaluated by burst pressure (BP) measurement. RESULTS: A fourfold increase in BP was observed in the SR+ HAM-PTB group compared to suture repair alone (94 ± 3 vs. 25 ± 8 mm Hg, P < 0.0001) at day 3. At day 7 the burst pressures were 165 ± 40 and 145 ± 31 mm Hg (P = 1), respectively. A significant decrease in peri-anastomotic adhesions was observed in the SR+ HAM-PTB group compared to the SR group at both time points (P < 0.001). CONCLUSION: Sealing sutured colonic anastomotic lines with HAM-PTB increases the early strength of the repair and reduces peri-anastomotic adhesions. Lasers Surg. Med. 48:530-537, 2016. © 2016 Wiley Periodicals, Inc.
- 6Tsao, S.; Yao, M.; Tsao, H.; Henry, F. P.; Zhao, Y.; Kochevar, J. J.; Redmond, R. W.; Kochevar, I. E. Light-activated tissue bonding for excisional wound closure: a split-lesion clinical trial Br. J. Dermatol. 2012, 166, 555– 563 DOI: 10.1111/j.1365-2133.2011.10710.xGoogle Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC383lsVCgug%253D%253D&md5=26e81b4b36431acc477589ade5ec3d80Light-activated tissue bonding for excisional wound closure: a split-lesion clinical trialTsao S; Yao M; Tsao H; Henry F P; Zhao Y; Kochevar J J; Redmond R W; Kochevar I EThe British journal of dermatology (2012), 166 (3), 555-63 ISSN:.BACKGROUND: Apposition of wound edges by sutures provides a temporary scaffold and tension support for healing. We have developed a novel tissue-sealing technology, photoactivated tissue bonding (PTB), which immediately crosslinks proteins between tissue planes, thereby sealing on a molecular scale. OBJECTIVES: To determine the effectiveness of PTB for superficial closure of skin excisions and to compare the results with standard epidermal suturing. METHODS: A split-lesion, paired comparison study of 31 skin excisions was performed. Following deep closure with absorbable sutures, one-half of each wound was superficially closed with nonabsorbable nylon sutures while the other half was stained with Rose Bengal dye and treated with green light. Overall appearance and scar characteristics were rated at 2weeks and 6months in a blinded manner by three dermatologists viewing photographs, by two onsite physicians and by patients. RESULTS: At 2weeks, neither sutured nor PTB-treated segments showed dehiscence; however, PTB-sealed segments showed less erythema than sutured segments as determined by photographic (P=0·001) and onsite evaluations (P=0·005). Overall appearance after PTB was judged better than after sutures (P=0·002). At 6months, scars produced by PTB were deemed superior to scars resulting from sutures in terms of appearance (P<0·001), width (P=0·002) and healing (P=0·003). Patients were more satisfied with the appearance of the PTB-sealed wound half after 2weeks and 6months (P=0·013 and P=0·003, respectively). CONCLUSIONS: A novel molecular suturing technique produces effective wound sealing and less scarring than closure with nylon interrupted epidermal sutures. Comparisons with better suturing techniques are warranted.
- 7Verter, E. E.; Gisel, T. E.; Yang, P.; Johnson, A. J.; Redmond, R. W.; Kochevar, I. E. Light-initiated bonding of amniotic membrane to cornea Invest. Ophthalmol. Visual Sci. 2011, 52, 9470– 9477 DOI: 10.1167/iovs.11-7248Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XjtVKkurY%253D&md5=2055c585b721cfd26530c542e03fd7a9Light-initiated bonding of amniotic membrane to corneaVerter, E. Eri; Gisel, Thomas E.; Yang, Penggao; Johnson, Anthony J.; Redmond, Robert W.; Kochevar, Irene E.Investigative Ophthalmology & Visual Science (2011), 52 (13), 9470-9477CODEN: IOVSDA; ISSN:1552-5783. (Association for Research in Vision and Ophthalmology)Purpose: Suturing amniotic membrane to cornea during surgery is time consuming, and sutures may further damage the eye. The authors introduce a novel sutureless, light-activated technique that securely attaches amnion to cornea through protein-protein crosslinks. Methods: Cryopreserved human amniotic membrane, stained with Rose Bengal (RB), was placed over a full-thickness wound in deepithelialized rabbit cornea and was treated with green laser. The intraocular pressure that broke the seal (IOPL) was measured, and adhesion was measured with a peel test. The influences on bonding strength of fluence, irradiance, RB concn., and amnion surface bonded were measured. Epithelial cell migration on treated amnion and keratocyte viability after bonding were also measured. The involvement in the bonding mechanism of oxygen, singlet oxygen, and assocn. of RB with stromal collagen was investigated. Results: Sealing amniotic membrane over cornea using 0.1% RB and 150 J/cm2 at 532 nm produced an IOPL of 261 ± 77 mm Hg ex vivo and 448 mm ± 212 mm Hg in vivo. The ex vivo IOPL increased with increasing fluence (50-150 J/cm2). Equivalent IOPL was produced for bonding basement membrane or stromal amnion surfaces. The bonding treatment was not toxic to keratocytes but slightly reduced the migration of corneal epithelial cells on amnion ex vivo. Mechanism studies indicated that RB forms two complexes with amnion stromal collagen, that bonding requires oxygen, and that singlet oxygen mediates protein crosslinking. Conclusions: A rapid, light-activated technique produces strong, immediate bonding between amnion and cornea and merits further evaluation for ocular surface surgeries.
- 8Cherfan, D.; Verter, E. E.; Melki, S.; Gisel, T. E.; Doyle, F. J., Jr.; Scarcelli, G.; Yun, S. H.; Redmond, R. W.; Kochevar, I. E. Collagen cross-linking using rose bengal and green light to increase corneal stiffness Invest. Ophthalmol. Visual Sci. 2013, 54, 3426– 3433 DOI: 10.1167/iovs.12-11509Google ScholarThere is no corresponding record for this reference.
- 9Zhu, H.; Alt, C.; Webb, R. H.; Melki, S.; Kochevar, I. E. Corneal Crosslinking With Rose Bengal and Green Light: Efficacy and Safety Evaluation Cornea 2016, 35, 1234– 1241 DOI: 10.1097/ICO.0000000000000916Google ScholarThere is no corresponding record for this reference.
- 10Goldstone, R. N.; McCormack, M. C.; Goldstein, R. L.; Mallidi, S.; Randolph, M. A.; Watkins, M. T.; Redmond, R. W.; Austen, W. G., Jr. Photochemical Tissue Passivation Attenuates AV Fistula Intimal HyperplasiaAnn. Surg. 2016, DOI: 10.1097/SLA.0000000000002046 .Google ScholarThere is no corresponding record for this reference.
- 11Goldstone, R. N.; McCormack, M. C.; Khan, S. I.; Salinas, H. M.; Meppelink, A.; Randolph, M. A.; Watkins, M. T.; Redmond, R. W.; Austen, W. G., Jr. Photochemical Tissue Passivation Reduces Vein Graft Intimal Hyperplasia in a Swine Model of Arteriovenous Bypass Grafting J. Am. Heart Assoc. 2016, 5e003856 DOI: 10.1161/JAHA.116.003856Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhvF2jsrjN&md5=55d0373eea312e7f41d626f8a28ecf39Photochemical tissue passivation reduces vein graft intimal hyperplasia in a swine model of arteriovenous bypass graftingGoldstone, Robert N.; McCormack, Michael C.; Khan, Saiqa I.; Salinas, Harry M.; Meppelink, Amanda; Randolph, Mark A.; Watkins, Michael T.; Redmond, Robert W.; Austen, William G. Jr.Journal of the American Heart Association (2016), 5 (8), e003856/1-e003856/10CODEN: JAHABZ; ISSN:2047-9980. (Wiley-Blackwell)Background-Bypass grafting remains the std. of care for coronary artery disease and severe lower extremity ischemia. Efficacy is limited by poor long-term venous graft patency secondary to intimal hyperplasia (IH) caused by venous injury upon exposure to arterial pressure. We investigate whether photochem. tissue passivation (PTP) treatment of vein grafts modulates smooth muscle cell (SMC) proliferation and migration, and inhibits development of IH. Methods and Results-PTP was performed at increasing fluences up to 120 J/cm2 on porcine veins. Tensiometry performed to assess vessel elasticity/stiffness showed increased stiffness with increasing fluence until plateauing at 90 J/cm2 (median, interquartile range [IQR]). At 90 J/cm2, PTP-treated vessels had a 10-fold greater Young's modulus than untreated controls (954 [IQR, 2217] vs 99 kPa [IQR, 63]; P=0.03). Each pig received a PTP-treated and untreated carotid artery venous interposition graft. At 4-wk, intimal/medial areas were assessed. PTP reduced the degree of IH by 66% and medial hypertrophy by 49%. Intimal area was 3.91 (IQR, 1.2) and 1.3 mm2 (IQR, 0.97; P≤0.001) in untreated and PTP-treated grafts, resp. Medial area was 9.2 (IQR, 3.2) and 4.7 mm2 (IQR, 2.0; P≤0.001) in untreated and PTP-treated grafts, resp. Immunohistochem. was performed to assess alpha-smooth muscle actin (SMA) and proliferating cell nuclear antigen (PCNA). Objectively, there were less SMA-pos. cells within the intima/media of PTP-treated vessels than controls. There was an increase in PCNA-pos. cells within control vein grafts (18% [IQR, 5.3]) vs. PTP-treated vein grafts (5% [IQR, 0.9]; P=0.02). Conclusions-By strengthening vein grafts, PTP decreases SMC proliferation and migration, thereby reducing IH.
- 12Fernandes, J. R.; Salinas, H. M.; Broelsch, G. F.; McCormack, M. C.; Meppelink, A. M.; Randolph, M. A.; Redmond, R. W.; Austen, W. G., Jr. Prevention of capsular contracture with photochemical tissue passivation Plast. Reconstr. Surg. 2014, 133, 571– 577 DOI: 10.1097/01.prs.0000438063.31043.79Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXjt1SntL0%253D&md5=a423afa38ba1d0cf77a9dab0853efb4dPrevention of Capsular Contracture with Photochemical Tissue PassivationFernandes, Justin R.; Salinas, Harry M.; Broelsch, G. Felix; McCormack, Michael C.; Meppelink, Amanda M.; Randolph, Mark A.; Redmond, Robert W.; Austen, William G., Jr.Plastic and Reconstructive Surgery (2014), 133 (3), 571-577CODEN: PRSUAS; ISSN:1529-4242. (Lippincott Williams & Wilkins)Background: Capsular contracture is the most common complication following the insertion of breast implants. Within a decade, half of patients will develop capsular contracture, leading to significant morbidity and need for reoperation. There is no preventative treatment available and the recurrence rate remains high. Photochem. tissue passivation is a novel tissue-stabilization technique that results in collagen crosslinking. It can rapidly link collagen fibers in situ, preserving normal tissue architecture. By using this therapy to passivate the collagenous tissues of the implant pocket, the authors hope to prevent the development of pathogenic collagen bundles and subsequent capsule contracture. Methods: Six-cubic centimeter tissue expanders were placed below the panniculus carnosus muscle along the dorsum of New Zealand white rabbits. Fibrin glue was instilled into each implant pocket to induce contracture. Treated pockets received photochem. tissue passivation by coating them with a photosensitizing dye and exposing the area to a 532-nm light. After 8 wk, capsule tissue was harvested for histol. evaluation. Results: Implant capsule thickness is the no. one prognostic factor for contracture development. The authors demonstrated a 52 percent decrease in capsule thickness in the passivated group compared with controls. Photochem. tissue passivation resulted in fewer fibrohistiocytic cells and macrophages and in reduced synovial metaplasia and smooth muscle actin deposition. Conclusions: Photochem. tissue passivation significantly decreased both capsule thickness and smooth muscle actin deposition. It is a promising technique for preventing capsular contracture that can be performed at the time of initial surgery without a significant increase in procedure time.
- 13Ludvikova, L.; Fris, P.; Heger, D.; Sebej, P.; Wirz, J.; Klan, P. Photochemistry of rose bengal in water and acetonitrile: a comprehensive kinetic analysis Phys. Chem. Chem. Phys. 2016, 18, 16266– 16273 DOI: 10.1039/C6CP01710JGoogle Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XptVCrurY%253D&md5=16827a5eff05f8af311b79d04e80a169Photochemistry of rose bengal in water and acetonitrile: a comprehensive kinetic analysisLudvikova, Lucie; Fris, Pavel; Heger, Dominik; Sebej, Peter; Wirz, Jakob; Klan, PetrPhysical Chemistry Chemical Physics (2016), 18 (24), 16266-16273CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The photophys. and photochem. properties of rose bengal (RB) in degassed aq. and acetonitrile solns. were studied using steady-state and transient absorption spectroscopies. This comprehensive investigation provides detailed information about the kinetics and the optical properties of all intermediates involved: the triplet excited state and the oxidized and reduced forms of RB. A full kinetic description is used to control the concns. of these intermediates by changing the initial exptl. conditions.
- 14Neckers, D. C. Rose Bengal J. Photochem. Photobiol. A 1989, 47, 1– 29 DOI: 10.1016/1010-6030(89)85002-6Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXktlWisbw%253D&md5=532dc3c2c508b43964c9247be4264613Rose BengalNeckers, D. C.Journal of Photochemistry and Photobiology, A: Chemistry (1989), 47 (1), 1-29CODEN: JPPCEJ; ISSN:1010-6030.A review with 38 refs. The spectral properties, photochem. reactivity, and photophys. parameters of all of the known derivs. of Rose Bengal, 2,4,5,7-tetraiodo-3',4',5',6'-tetrachlorofluorescein, are reported and compared.
- 15Fleming, G. R.; Morris, J. M.; Morrison, R. J. S.; Robinson, G. W. Picosecond fluorescence studies of xanthene dyes J. Am. Chem. Soc. 1977, 99, 4306– 4311 DOI: 10.1021/ja00455a017Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2sXkvVCjsrk%253D&md5=d2ae45a7fa0caf469a6dcf0b5cc9a0e1Picosecond fluorescence studies of xanthene dyesFleming, G. R.; Knight, A. W. E.; Morris, J. M.; Morrison, R. J. S.; Robinson, G. W.Journal of the American Chemical Society (1977), 99 (13), 4306-11CODEN: JACSAT; ISSN:0002-7863.Subnanosecond lifetime measurements using picosecond pulses from a mode-locked Nd3+-glass laser together with conventional absorption and fluorescence-yield methods are used to study the photophys. of fluorescein, eosin, erythrosin, and rose bengal in aq. and alc. solvents. For each of the dye mols., absorption and fluorescence max. move towards higher energy (blue shift) as the solvent changes from Me2CHOH to H2O. Fluorescence lifetimes and quantum yields decrease with this solvent change and also with increased halogenation of the fluorescein parent, owing to the heavy-atom effect. The variations obsd. in the nonradiative part of the decay rate are attributed to variations in the rate of S1-T1 intersystem crossing. For these particular solvent-solute combinations stabilization energies lie in the order ΔE(T1) < ΔE(S1) < ΔE(S0). This is consistent with both the increased S1-S0 spectral blue shifts and the enhanced intersystem crossing rate, arising from a smaller S1-T1 energy gap, when these dye mols. are placed in a more aq. solvent environment. The use of these dyes as fluorescent probes in biol. important mols. is discussed.
- 16Kochevar, I. E.; Redmond, R. W. Photosensitized production of singlet oxygen Methods Enzymol. 2000, 319, 20– 28 DOI: 10.1016/s0076-6879(00)19004-4Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXlsl2lsbo%253D&md5=46bb84b5807a1fe38a6afa20fc977996Photosensitized production of singlet oxygenKochevar, Irene E.; Redmond, Robert W.Methods in Enzymology (2000), 319 (), 20-28CODEN: MENZAU; ISSN:0076-6879. (Academic Press)A review with 8 refs. Photosensitization is a simple and controllable method for the generation of singlet oxygen in soln. and in cells. Methods are described for detg. the yield of singlet oxygen in soln., for measurement of the rate of reaction between single oxygen and a substrate, and for comparing the effectiveness of singlet oxygen generated by different photosensitizers in cells. These quant. measurements can lead to better understanding of the interaction of singlet oxygen with biomols. (c) 2000 Academic Press.
- 17Shen, H. R.; Spikes, J. D.; Kopeckova, P.; Kopecek, J. Photodynamic crosslinking of proteins. II. Photocrosslinking of a model protein-ribonuclease A J. Photochem. Photobiol. B 1996, 35, 213– 219 DOI: 10.1016/S1011-1344(96)07300-9Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xms1emurk%253D&md5=b2f68b4acb7c761ec67244bdf0eab706Photodynamic crosslinking of proteins. II. Photocrosslinking of a model protein-ribonuclease AShen, Hui-Rong; Spikes, John D.; Kopeckova, Pavla; Kopecek, JindrichJournal of Photochemistry and Photobiology, B: Biology (1996), 35 (3), 213-219CODEN: JPPBEG; ISSN:1011-1344. (Elsevier)Illumination of bovine pancreatic RNase A (RNase A) in soln. in the presence of rose bengal as a photosensitizer resulted in the progressive formation of enzyme dimers, trimers, tetramers and higher oligomers, as measured by gel electrophoresis and size exclusion chromatog. Oxygen was necessary for crosslink formation, and azide inhibition studies indicated that singlet oxygen was involved in the process. Chem. modification of His residues (with di-Et pyrocarbonate) and/or Lys residues (with acetic acid N-hydroxysuccinimide ester) in the enzyme decreased crosslinking, suggesting the participation of these two amino acid residues in the reaction. Met and cystine residues did not appear to be involved. Similar studies have shown that model N-(2-hydroxypropyl)methacrylamide (HPMA) copolymers contg. ε-aminocaproic acid side chains terminating in His or Lys residues are photodynamically crosslinked via His-His or His-Lys interactions. Treatment of crosslinked RNase A and its His, Lys and Lys-His derivs. for 5 min at 97 °C in a dithiothreitol-sodium dodecyl sulfate mixt. efficiently ruptured a major part of the photodynamically formed crosslinks; treatment with the detergent alone had no effect. Similar results were previously obtained with the crosslinked amino acid-contg. HPMA copolymers, suggesting that photodynamic crosslinks involving His-His and His-Lys interaction are chem. the same in RNase A and the copolymer model.
- 18Lambert, C. R.; Kochevar, I. E. Electron transfer quenching of the rose bengal triplet state Photochem. Photobiol. 1997, 66, 15– 25 DOI: 10.1111/j.1751-1097.1997.tb03133.xGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXksF2ksbw%253D&md5=d9f0db65a5fe55e35ea986835222e958Electron transfer quenching of the rose bengal triplet stateLambert, Christopher R.; Kochevar, Irene E.Photochemistry and Photobiology (1997), 66 (1), 15-25CODEN: PHCBAP; ISSN:0031-8655. (American Society for Photobiology)The potential for electron transfer quenching of rose bengal triplet (3RB2-) to compete with energy transfer quenching by oxygen was evaluated. Rate consts. for oxidative and reductive quenching were measured in buffered aq. soln., acetonitrile and in small unilamellar liposomes using laser flash photolysis. Biol. relevant quenchers were used that varied widely in structure, redn. potential and charge. Radical ion yields (φi) were measured by monitoring the absorption of the rose bengal semireduced (RB·3-) and semioxidized (RB·-) radicals. The results in soln. were analyzed as a function of the free energy for electron transfer (ΔG) calcd. using the Weller equation including electrostatic terms. Exothermic oxidative quenching was about 10-fold faster than exothermic reductive quenching in aq. soln. The quenching rate consts. decreased as ΔG approached zero in both aq. and acetonitrile soln. Exceptions to these generalizations were obsd. that could be rationalized by specific steric or electrostatic effects or by a change in mechanism. The results suggest that electron transfer reactions with some potential quenchers in cells could compete with formation of singlet oxygen [O2(1Δg)]. Values of φi were generally greater for reductive quenching and, for oxidative quenching, greater in acetonitrile than in buffer. Electron transfer quenching of 3RB2- in liposomes, below the phase transition temp. was slower than in soln. for both lipid-sol. and water-sol. quenchers indicating that these reactions may not compete with formation of (O21Δg) during cell photosensitization.
- 19Zakrzewski, A.; Neckers, D. C. Bleaching products of rose bengal under reducing conditions Tetrahedron 1987, 43, 4507– 4512 DOI: 10.1016/S0040-4020(01)86891-5Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXkslaku70%253D&md5=e1bb13110d942833b5df2b176055e93eBleaching products of Rose Bengal under reducing conditionsZakrzewski, Andrzej; Neckers, D. C.Tetrahedron (1987), 43 (20), 4507-12CODEN: TETRAB; ISSN:0040-4020.The bleaching behavior of Rose Bengal under reducing conditions was elucidated by detg. the products of chem. and photochem. redn. of Rose Bengal.
- 20Oster, G.; Oster, G. K.; Karg, G. Extremely long-lived intermediates in photochemical reactions of dyes in non-visous media J. Phys. Chem. 1962, 66, 2514– 2517 DOI: 10.1021/j100818a045Google ScholarThere is no corresponding record for this reference.
- 21Luttrull, D. K.; Valdes-Aguilera, O.; Linden, S. M.; Paczkowski, J.; Neckers, D. C. Rose Bengal aggregation in rationally synthesized dimeric systems Photochem. Photobiol. 1988, 47, 551– 557 DOI: 10.1111/j.1751-1097.1988.tb08843.xGoogle ScholarThere is no corresponding record for this reference.
- 22Simpson, M. J.; Poblete, H.; Griffith, M.; Alarcon, E. I.; Scaiano, J. C. Impact of dye-protein interaction and silver nanoparticles on rose bengal photophysical behavior and protein photocrosslinking Photochem. Photobiol. 2013, 89, 1433– 1441 DOI: 10.1111/php.12119Google ScholarThere is no corresponding record for this reference.
- 23Valdes-Aguilera, O.; Neckers, D. C. Aggregation phenomena in xanthene dyes Acc. Chem. Res. 1989, 22, 171– 177 DOI: 10.1021/ar00161a002Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXitFCksb8%253D&md5=2feeb8679bb8c7cf735a59e3aa19590aAggregation phenomena in xanthene dyesValdes-Aguilera, Oscar; Neckers, D. C.Accounts of Chemical Research (1989), 22 (5), 171-7CODEN: ACHRE4; ISSN:0001-4842.A crit. review with 63 refs. on aggregation of xanthene dyes, principally rhodamines and fluoresceins. The author's work on Rose Bengal is considered in detail.
- 24Xu, D.; Neckers, D. C. Aggregation of rose bengal molecules in solution J. Photochem. Photobiol. A 1987, 40, 361– 370 DOI: 10.1016/1010-6030(87)85013-XGoogle Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXmtlymtL0%253D&md5=239ea45bfc0643e238061f40c67e6130Aggregation of Rose Bengal molecules in solutionXu, Danian; Neckers, D. C.Journal of Photochemistry and Photobiology, A: Chemistry (1987), 40 (2-3), 361-70CODEN: JPPCEJ; ISSN:1010-6030.Absorption, emission, and excitation spectral data support the hypothesis that Rose Bengal forms H-type aggregates in water and in polar, protic solvents.
- 25Alarcón, E.; Edwards, A. M.; Aspee, A.; Borsarelli, C. D.; Lissi, E. A. Photophysics and photochemistry of rose bengal bound to human serum albumin Photochem. Photobiol. Sci. 2009, 8, 933– 943 DOI: 10.1039/b901056dGoogle Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXotVenu7s%253D&md5=c520abf9b3cdc4448c16998478eafef5Photophysics and photochemistry of rose bengal bound to human serum albuminAlarcon, Emilio; Edwards, Ana Maria; Aspee, Alexis; Borsarelli, Claudio D.; Lissi, Eduardo A.Photochemical & Photobiological Sciences (2009), 8 (7), 933-943CODEN: PPSHCB; ISSN:1474-905X. (Royal Society of Chemistry)Rose bengal (RB) readily binds to human serum albumin (HSA). At low RB concns., 90% of the dye is assocd. to the protein (5 μM). This assocn. takes place in specific binding sites I and/or II. At higher RB concns., unspecific binding takes place with up to 10 RB mols. bound per protein mol. The behavior of excited RB mols. bound to HSA is widely different to that obsd. in aq. soln. Furthermore, the data also show that the behavior of bound RB mols. changes with the av. no. of dye mols. per protein (n). In particular, when n is large, the fluorescence yield is significantly reduced and no measurable long-lived triples and free singlet oxygen formation from bound dyes is detected. These results are related to self-quenching of the singlet and, most likely, excited triplets. All results point to the relevance of intra-protein generated singlet oxygen. However, when the dye is bound to the protein, at low oxygen concns. such as those prevailing in vivo, trapping by oxygen of the triplet becomes inefficient and type I processes could contribute to the obsd. photoprocesses.
- 26Ni, T.; Senthil-Kumar, P.; Dubbin, K.; Aznar-Cervantes, S. D.; Datta, N.; Randolph, M. A.; Cenis, J. L.; Rutledge, G. C.; Kochevar, I. E.; Redmond, R. W. A photoactivated nanofiber graft material for augmented Achilles tendon repair Lasers Surg. Med. 2012, 44, 645– 652 DOI: 10.1002/lsm.22066Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC38bgt1WhtQ%253D%253D&md5=f59c0a45f8df812d81e675aa7e0a2011A photoactivated nanofiber graft material for augmented Achilles tendon repairNi Tao; Senthil-Kumar Prabhu; Dubbin Karen; Aznar-Cervantes Salvador D; Datta Neha; Randolph Mark A; Cenis Jose L; Rutledge Gregory C; Kochevar Irene E; Redmond Robert WLasers in surgery and medicine (2012), 44 (8), 645-52 ISSN:.BACKGROUND AND OBJECTIVE: Suture repair of Achilles tendon rupture can cause infection, inflammation and scarring, while prolonged immobilization promotes adhesions to surrounding tissues and joint stiffness. Early mobilization can reduce complications provided the repair is strong enough to resist re-rupture. We have developed a biocompatible, photoactivated tendon wrap from electrospun silk (ES) to provide additional strength to the repair that could permit early mobilization, and act as a barrier to adhesion formation. STUDY DESIGN/MATERIAL AND METHODS: ES nanofiber mats were prepared by electrospinning. New Zealand white rabbits underwent surgical transection of the Achilles tendon and repair by: (a) SR: standard Kessler suture + epitendinous suture (5-0 vicryl). (b) ES/PTB: a single stay suture and a section of ES mat, stained with 0.1% Rose Bengal (RB), wrapped around the tendon and bonded with 532 nm light (0.3 W/cm(2) , 125 J/cm(2) ). (c) SR + ES/PTB: a combination of (a) and (b). Gross appearance, extent of adhesion formation and biomechanical properties of the repaired tendon were evaluated at Days 7, 14, or 28 post-operatively (n = 8 per group at each time point). RESULTS: Ultimate stress (US) and Young's modulus (E) in the SR group were not significantly different from the ES/PTB group at Days 7 (US, P = 0.85; E, P = 1), 14 (US, P = 0.054; E, P = 1), and 28 (US, P = 0.198; E, P = 0.12) post-operatively. Adhesions were considerably greater in the SR group compared to the ES/PTB group at Days 7 (P = 0.002), 14 (P < 0.0001), and 28 (P < 0.0001). The combination approach of SR + ES/PTB gave the best outcomes in terms of E at 7 (P < 0.016) and 14 days (P < 0.016) and reduced adhesions compared to SR at 7 (P < 0.0001) and 14 days (P < 0.0001), the latter suggesting a barrier function for the photobonded ES wrap. CONCLUSION: Photochemical sealing of a ES mat around the tendon repair site provides considerable benefit in Achilles tendon repair. Lasers Surg. Med. 44: 645-652, 2012. © 2012 Wiley Periodicals, Inc.
- 27Stetefeld, J.; Frank, S.; Jenny, M.; Schulthess, T.; Kammerer, R. A.; Boudko, S.; Landwehr, R.; Okuyama, K.; Engel, J. Collagen stabilization at atomic level: crystal structure of designed (GlyProPro)10foldon Structure 2003, 11, 339– 346 DOI: 10.1016/S0969-2126(03)00025-XGoogle Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXhvFait7g%253D&md5=75307e701596341cc24da4e3ae1bd7c1Collagen Stabilization at Atomic Level Crystal Structure of Designed (GlyProPro)10foldonStetefeld, Jorg; Frank, Sabine; Jenny, Margrit; Schulthess, Therese; Kammerer, Richard A.; Boudko, Sergei; Landwehr, Ruth; Okuyama, Kenji; Engel, JurgenStructure (Cambridge, MA, United States) (2003), 11 (3), 339-346CODEN: STRUE6; ISSN:0969-2126. (Cell Press)In a designed fusion protein the trimeric domain foldon from bacteriophage T4 fibritin was connected to the C terminus of the collagen model peptide (GlyProPro)10 by a short Gly-Ser linker to facilitate formation of the three-stranded collagen triple helix. Crystal structure anal. at 2.6 A resoln. revealed conformational changes within the interface of both domains compared with the structure of the isolated mols. A striking feature is an angle of 62.5° between the symmetry axis of the foldon trimer and the axis of the triple helix. The melting temp. of (GlyProPro)10 in the designed fusion protein (GlyProPro)10foldon is higher than that of isolated (GlyProPro)10, which suggests an entropic stabilization compensating for the destabilization at the interface.
- 28Mirenda, M.; Dicelio, L. E.; San Roman, E. Effect of molecular interactions on the photophysics of Rose Bengal in polyelectrolyte solutions and self-assembled thin films J. Phys. Chem. B 2008, 112, 12201– 12207 DOI: 10.1021/jp803892gGoogle Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhtVymtLrK&md5=c6858f9042e0b1abc53c70d13981c743Effect of molecular interactions on the photophysics of Rose Bengal in polyelectrolyte solutions and self-assembled thin filmsMirenda, Martin; Dicelio, Lelia E.; San Roman, EnriqueJournal of Physical Chemistry B (2008), 112 (39), 12201-12207CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Aq. solns. and layer-by-layer self-assembled thin films contg. Rose Bengal and poly(diallyldimethylammonium chloride) are studied with the aim of understanding the interactions controlling their structures and the photophysics of the dye in both media. A detailed spectroscopic and theor. anal. shows that hydrophobic interactions among dye mols. contribute to the coiling of the polyelectrolyte chain in soln. at low polyelectrolyte/dye (P/D) ratios, whereas extensive aggregation of the dye takes place even at ratios as high as 104 (expressed in monomeric units). A polyelectrolyte elongated form prevails in self-assembled thin films, providing an environment that reduces hydrophobic interactions and lowers the aggregation tendency. Self-assembled films with a roughly estd. overall dye concn. around 1 M at a P/D ratio in the order of seven are fluorescent and photogenerate singlet mol. oxygen. This contrasts with the behavior of polyelectrolyte solns., which are almost nonfluorescent and do not evidence triplet state generation at the same P/D ratio.
- 29Rodríguez, H. B.; Lagorio, M. G.; San Roman, E. Rose Bengal adsorbed on microgranular cellulose: evidence on fluorescent dimers Photochem. Photobiol. Sci. 2004, 3, 674– 680 DOI: 10.1039/B402484BGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXltlynu7Y%253D&md5=36a73d53cdc4a45b8af54f12e88ad5feRose Bengal adsorbed on microgranular cellulose: evidence on fluorescent dimersRodriguez, Hernan B.; Lagorio, M. Gabriela; San Roman, EnriquePhotochemical & Photobiological Sciences (2004), 3 (7), 674-680CODEN: PPSHCB; ISSN:1474-905X. (Royal Society of Chemistry)Rose Bengal adsorbed on microgranular cellulose was studied in the solid phase by total and diffuse reflectance and steady-state emission spectroscopy. A simple self-assocn. monomer-dimer equil. fitted reflectance data up to dye loadings of 4 × 10-7 mol (g cellulose)-1 and allowed calcn. of monomer and dimer spectra. Further increase of dye loading resulted in the formation of higher aggregates. Obsd. emission and excitation spectra and quantum yields were cor. for reabsorption and reemission of luminescence, using a previously developed model, within the assumption that only monomers are luminescent. An apparent increase of fluorescence quantum yield with dye loading was found, which was attributed to the occurrence of dimer fluorescence. Extension of the model to two luminescent species (i.e. monomer and dimer) yielded const. fluorescence quantum yields for the monomer, ΦM = 0.120 ± 0.004, and for the dimer, ΦD = 0.070 ± 0.006. The monomer quantum yield is close to the value found for the same dye in basic ethanol. The presence of fluorescent dimers and calcd. quantum yields are supported by anal. of the excitation spectra and other exptl. evidence. The possible occurrence of non-radiative energy transfer and the effect of surface charge on the properties of the dimer are analyzed.
- 30Lissi, E. A.; Encinas, M. V.; Lemp, E.; Rubio, M. A. Singlet oxygen O2(1.DELTA.g) bimolecular processes. Solvent and compartmentalization effects Chem. Rev. 1993, 93, 699– 723 DOI: 10.1021/cr00018a004Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXhvFSrt7Y%253D&md5=186df1c5c33fc327b00db39d8190c2efSinglet oxygen O2(1Δg) bimolecular processes. Solvent and compartmentalization effectsLissi, E. A.; Encinas, M. V.; Lemp, E.; Rubio, M. A.Chemical Reviews (Washington, DC, United States) (1993), 93 (2), 699-723CODEN: CHREAY; ISSN:0009-2665.A review with >401 refs. including (1) deactivation dominated by energy transfer, (2) dyes and sensitizers, (3) inorg. anions, (4) N-contg. compds., (5) S-contg. compds., (6) unsatd. compds., and (7) microheterogeneous systems.
- 31Alarcón, E.; Edwards, A. M.; Aspee, A.; Moran, F. E.; Borsarelli, C. D.; Lissi, E. A.; Gonzalez-Nilo, D.; Poblete, H.; Scaiano, J. C. Photophysics and photochemistry of dyes bound to human serum albumin are determined by the dye localization Photochem. Photobiol. Sci. 2010, 9, 93– 102 DOI: 10.1039/B9PP00091GGoogle Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXht1ehsA%253D%253D&md5=ef562f3d70b318e169d5d987e421859fPhotophysics and photochemistry of dyes bound to human serum albumin are determined by the dye localizationAlarcon, Emilio; Edwards, Ana Maria; Aspee, Alexis; Moran, Faustino E.; Borsarelli, Claudio D.; Lissi, Eduardo A.; Gonzalez-Nilo, Danilo; Poblete, Horacio; Scaiano, J. C.Photochemical & Photobiological Sciences (2010), 9 (1), 93-102CODEN: PPSHCB; ISSN:1474-905X. (Royal Society of Chemistry)The photophysics and photochem. of rose bengal (RB) and methylene blue (MB) bound to human serum albumin (HSA) have been investigated under a variety of exptl. conditions. Distribution of the dyes between the external solvent and the protein has been estd. by phys. sepn. and fluorescence measurements. The main localization of protein-bound dye mols. was estd. by the intrinsic fluorescence quenching, displacement of fluorescent probes bound to specific protein sites, and by docking modeling. All the data indicate that, at low occupation nos., RB binds strongly to the HSA site I, while MB localizes predominantly in the protein binding site II. This different localization explains the obsd. differences in the dyes' photochem. behavior. In particular, the environment provided by site I is less polar and considerably less accessible to oxygen. The localization of RB in site I also leads to an efficient quenching of the intrinsic protein fluorescence (ascribed to the nearby Trp residue) and the generation of intra-protein singlet oxygen, whose behavior is different to that obsd. in the external solvent or when it is generated by bound MB.
- 32Mills, A.; Lawrence, C.; Douglas, P. Photoreduction of Water sensitised by Rose Bengal J. Chem. Soc., Faraday Trans. 2 1986, 82, 2291– 2303 DOI: 10.1039/f29868202291Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2sXisFKhtg%253D%253D&md5=61fb2cba0514241a10dc63b81516dc44Photoreduction of water sensitized by Rose BengalMills, Andrew; Lawrence, Carl; Douglas, PeterJournal of the Chemical Society, Faraday Transactions 2: Molecular and Chemical Physics (1986), 82 (12), 2291-303CODEN: JCFTBS; ISSN:0300-9238.Rose Bengal was used to sensitize the photoredn. of H2O in aq. EtOH soln. (5% EtOH vol./vol.; ph 4.4) via both a reductive and an oxidative quenching mechanism, using EDTA and methylviologen, resp., as quenchers of its triplet state. In the presence of EDTA (which acts as both triplet quencher and sacrificial electron donor) the photoredn. of H2O sensitized by Rose Bengal was inefficient for a no. of reasons, including a low triplet quenching rate const. (8 × 103 dm3 mol-1 s-1) and a rapid disproportionation reaction involving the protonated semi-reduced dye radicals. In the presence of methylviologen (triplet quencher) and EDTA (sacrificial electron donor) the photoredn. of H2O sensitized by Rose Bengal was inefficient due to a low cage-escape yield (.vphi. ≈ 5 × 10-2) following the primary electron-transfer step, a fast back reaction (k ≥ 1 × 108 dm3 mol-1 s-1) and formation of a complex between the dye and methylviologen (K = 6400 ± 300 cm3 mol-). The complexed form of Rose Bengal appeared able to photosensitize, inefficiently, the photoredn. of methylviologen in the presence of EDTA, without exhibiting problems of dye fade.
- 33Pupkaite, J.; Ahumada, M.; McLaughlin, S.; Temkit, M.; Alaziz, S.; Seymour, R.; Ruel, M.; Kochevar, I.; Griffith, M.; Suuronen, E. J.; Alarcon, E. I. Collagen-Based Photoactive Agent for Tissue Bonding ACS Appl. Mater. Interfaces 2017, 9, 9265– 9270 DOI: 10.1021/acsami.7b01984Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXktVSitrg%253D&md5=b64c02792565822c5f6ac13db97e9755Collagen-Based Photoactive Agent for Tissue BondingPupkaite, Justina; Ahumada, Manuel; Mclaughlin, Sarah; Temkit, Maha; Alaziz, Sura; Seymour, Richard; Ruel, Marc; Kochevar, Irene; Griffith, May; Suuronen, Erik J.; Alarcon, Emilio I.ACS Applied Materials & Interfaces (2017), 9 (11), 9265-9270CODEN: AAMICK; ISSN:1944-8244. (American Chemical Society)Using a combination of methacrylated collagen and the photosensitizer rose Bengal, a new light-activated biomimetic material for tissue sutureless bonding was developed. This formulation was crosslinked using green light. In vivo tests in mice demonstrate the suitability of the material for sutureless wound closure.
- 34Henderson, B. W.; Busch, T. M.; Vaughan, L. A.; Frawley, N. P.; Babich, D.; Sosa, T. A.; Zollo, J. D.; Dee, A. S.; Cooper, M. T.; Bellnier, D. A.; Greco, W. R.; Oseroff, A. R. Photofrin photodynamic therapy can significantly deplete or preserve oxygenation in human basal cell carcinomas during treatment, depending on fluence rate Cancer Res. 2000, 60, 525– 529Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXht1Gmtrc%253D&md5=54600117edbfbf5e2b53454a37acc235Photofrin photodynamic therapy can significantly deplete or preserve oxygenation in human basal cell carcinomas during treatment, depending on fluence rateHenderson, Barbara W.; Busch, Theresa M.; Vaughan, Lurine A.; Frawley, Noreen P.; Babich, Debra; Sosa, Tara A.; Zollo, Joseph D.; Dee, Anthony S.; Cooper, Michele T.; Bellnier, David A.; Greco, William R.; Oseroff, Allan R.Cancer Research (2000), 60 (3), 525-529CODEN: CNREA8; ISSN:0008-5472. (AACR Subscription Office)At high fluence rates in animal models, photodynamic therapy (PDT) can photochem. deplete ambient tumor oxygen through the generation of singlet oxygen, causing acute hypoxia and limiting treatment effectiveness. We report that std. clin. treatment conditions (1 mg/kg Photofrin, light at 630 nm and 150 mW/cm2), which are highly effective for treating human basal cell carcinomas, significantly diminished tumor oxygen levels during initial light delivery in a majority of carcinomas. Oxygen depletion could be found during at least 40% of the total light dose, but tumors appeared well oxygenated toward the end of treatment. In contrast, initial light delivery at a lower fluence rate of 30 mW/cm2 increased tumor oxygenation in a majority of carcinomas. Laser treatment caused an intensity- and treatment time-dependent increase in tumor temp. The data suggest that high fluence rate treatment, although effective, may be inefficient.
- 35Ravichandran, R.; Islam, M. M.; Alarcon, E. I.; Samanta, A.; Wang, S.; Lundstrom, P.; Hilborn, J.; Griffith, M.; Phopase, J. Functionalised type-I collagen as a hydrogel building block for bio-orthogonal tissue engineering applications J. Mater. Chem. B. 2016, 4, 318– 326 DOI: 10.1039/C5TB02035BGoogle Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvVyqsb3M&md5=b58d7b77544ce08f78d4e548e7b32e6bFunctionalised type-I collagen as a hydrogel building block for bio-orthogonal tissue engineering applicationsRavichandran, R.; Islam, M. M.; Alarcon, E. I.; Samanta, A.; Wang, S.; Lundstroem, P.; Hilborn, J.; Griffith, M.; Phopase, J.Journal of Materials Chemistry B: Materials for Biology and Medicine (2016), 4 (2), 318-326CODEN: JMCBDV; ISSN:2050-7518. (Royal Society of Chemistry)In this study, we derivatized type I collagen without altering its triple helical conformation to allow for facile hydrogel formation via the Michael addn. of thiols to methacrylates without the addn. of other crosslinking agents. This method provides the flexibility needed for the fabrication of injectable hydrogels or pre-fabricated implantable scaffolds, using the same components by tuning the modulus from Pa to kPa. Enzymic degradability of the hydrogels can also be easily fine-tuned by variation of the ratio and the type of the crosslinking component. The structural morphol. reveals a lamellar structure mimicking native collagen fibrils. The versatility of this material is demonstrated by its use as a pre-fabricated substrate for culturing human corneal epithelial cells and as an injectable hydrogel for 3-D encapsulation of cardiac progenitor cells.
- 36Biasini, M.; Bienert, S.; Waterhouse, A.; Arnold, K.; Studer, G.; Schmidt, T.; Kiefer, F.; Gallo Cassarino, T.; Bertoni, M.; Bordoli, L.; Schwede, T. SWISS-MODEL: modelling protein tertiary and quaternary structure using evolutionary information Nucleic Acids Res. 2014, 42, W252– W258 DOI: 10.1093/nar/gku340Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFCqs73I&md5=51509ba8353b06f286d954ebe7e6673aSWISS-MODEL: modelling protein tertiary and quaternary structure using evolutionary informationBiasini, Marco; Bienert, Stefan; Waterhouse, Andrew; Arnold, Konstantin; Studer, Gabriel; Schmidt, Tobias; Kiefer, Florian; Cassarino, Tiziano Gallo; Bertoni, Martino; Bordoli, Lorenza; Schwede, TorstenNucleic Acids Research (2014), 42 (W1), W252-W258CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)Protein structure homol. modeling has become a routine technique to generate 3D models for proteins when exptl. structures are not available. Fully automated servers such as SWISS-MODEL with user-friendly web interfaces generate reliable models without the need for complex software packages or downloading large databases. Here, we describe the latest version of the SWISS-MODEL expert system for protein structure modeling. The SWISS-MODEL template library provides annotation of quaternary structure and essential ligands and co-factors to allow for building of complete structural models, including their oligomeric structure. The improved SWISS-MODEL pipeline makes extensive use of model quality estn. for selection of the most suitable templates and provides ests. of the expected accuracy of the resulting models. The accuracy of the models generated by SWISS-MODEL is continuously evaluated by the CAMEO system. The new web site allows users to interactively search for templates, cluster them by sequence similarity, structurally compare alternative templates and select the ones to be used for model building. In cases where multiple alternative template structures are available for a protein of interest, a user-guided template selection step allows building models in different functional states. SWISS-MODEL is available at http://swissmodel.expasy.org/.
- 37Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr. CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields J. Comput. Chem. 2010, 31, 671– 690 DOI: 10.1002/jcc.21367Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhtlentbc%253D&md5=26212e0e4f73bded0c89d4b411cd3833CHARMM general force field: A force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fieldsVanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; Mackerell, A. D., Jr.Journal of Computational Chemistry (2010), 31 (4), 671-690CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)The widely used CHARMM additive all-atom force field includes parameters for proteins, nucleic acids, lipids, and carbohydrates. In the present article, an extension of the CHARMM force field to drug-like mols. is presented. The resulting CHARMM General Force Field (CGenFF) covers a wide range of chem. groups present in biomols. and drug-like mols., including a large no. of heterocyclic scaffolds. The parametrization philosophy behind the force field focuses on quality at the expense of transferability, with the implementation concg. on an extensible force field. Statistics related to the quality of the parametrization with a focus on exptl. validation are presented. Addnl., the parametrization procedure, described fully in the present article in the context of the model systems, pyrrolidine, and 3-phenoxymethyl-pyrrolidine will allow users to readily extend the force field to chem. groups that are not explicitly covered in the force field as well as add functional groups to and link together mols. already available in the force field. CGenFF thus makes it possible to perform "all-CHARMM" simulations on drug-target interactions thereby extending the utility of CHARMM force fields to medicinally relevant systems. © 2009 Wiley Periodicals, Inc.J Comput Chem, 2010.
- 38Vanommeslaeghe, K.; MacKerell, A. D., Jr. Automation of the CHARMM General Force Field (CGenFF) I: bond perception and atom typing J. Chem. Inf. Model. 2012, 52, 3144– 3154 DOI: 10.1021/ci300363cGoogle Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fL&md5=c6679293f4a2501f2bcadf2020ca1473Automation of the CHARMM General Force Field (CGenFF) I: Bond Perception and Atom TypingVanommeslaeghe, K.; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3144-3154CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug-like mols. alone or interacting with biol. systems. In simulations involving biol. macromols., the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and charges is required. In the present article, which is part I of a series of two, we present the algorithms for bond perception and atom typing for the CHARMM General Force Field (CGenFF). The CGenFF atom typer first assocs. attributes to the atoms and bonds in a mol., such as valence, bond order, and ring membership among others. Of note are a no. of features that are specifically required for CGenFF. This information is then used by the atom typing routine to assign CGenFF atom types based on a programmable decision tree. This allows for straight-forward implementation of CGenFF's complicated atom typing rules and for equally straight-forward updating of the atom typing scheme as the force field grows. The presented atom typer was validated by assigning correct atom types on 477 model compds. including in the training set as well as 126 test-set mols. that were constructed to specifically verify its different components. The program may be utilized via an online implementation at https://www.paramchem.org/.
- 39Vanommeslaeghe, K.; Raman, E. P.; MacKerell, A. D., Jr. Automation of the CHARMM General Force Field (CGenFF) II: assignment of bonded parameters and partial atomic charges J. Chem. Inf. Model. 2012, 52, 3155– 3168 DOI: 10.1021/ci3003649Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Gns7fF&md5=e676ad1f42cb1e98dd353d4d285e8d13Automation of the CHARMM General Force Field (CGenFF) II: Assignment of Bonded Parameters and Partial Atomic ChargesVanommeslaeghe, K.; Raman, E. Prabhu; MacKerell, A. D.Journal of Chemical Information and Modeling (2012), 52 (12), 3155-3168CODEN: JCISD8; ISSN:1549-9596. (American Chemical Society)Mol. mechanics force fields are widely used in computer-aided drug design for the study of drug candidates interacting with biol. systems. In these simulations, the biol. part is typically represented by a specialized biomol. force field, while the drug is represented by a matching general (org.) force field. In order to apply these general force fields to an arbitrary drug-like mol., functionality for assignment of atom types, parameters, and partial at. charges is required. In the present article, algorithms for the assignment of parameters and charges for the CHARMM General Force Field (CGenFF) are presented. These algorithms rely on the existing parameters and charges that were detd. as part of the parametrization of the force field. Bonded parameters are assigned based on the similarity between the atom types that define said parameters, while charges are detd. using an extended bond-charge increment scheme. Charge increments were optimized to reproduce the charges on model compds. that were part of the parametrization of the force field. Case studies are presented to clarify the functioning of the algorithms and the significance of their output data.
- 40Best, R. B.; Zhu, X.; Shim, J.; Lopes, P. E.; Mittal, J.; Feig, M.; Mackerell, A. D., Jr. Optimization of the additive CHARMM all-atom protein force field targeting improved sampling of the backbone phi, psi and side-chain chi(1) and chi(2) dihedral angles J. Chem. Theory. Comput. 2012, 8, 3257– 3273 DOI: 10.1021/ct300400xGoogle Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtVKqurfP&md5=9a48a0c5770fb1e887c3bb34d45b1354Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone .vphi., ψ and Side-Chain χ1 and χ2 Dihedral AnglesBest, Robert B.; Zhu, Xiao; Shim, Jihyun; Lopes, Pedro E. M.; Mittal, Jeetain; Feig, Michael; MacKerell, Alexander D.Journal of Chemical Theory and Computation (2012), 8 (9), 3257-3273CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)While the quality of the current CHARMM22/CMAP additive force field for proteins has been demonstrated in a large no. of applications, limitations in the model with respect to the equil. between the sampling of helical and extended conformations in folding simulations have been noted. To overcome this, as well as make other improvements in the model, we present a combination of refinements that should result in enhanced accuracy in simulations of proteins. The common (non-Gly, -Pro) backbone CMAP potential has been refined against exptl. soln. NMR data for weakly structured peptides, resulting in a rebalancing of the energies of the α-helix and extended regions of the Ramachandran map, correcting the α-helical bias of CHARMM22/CMAP. The Gly and Pro CMAPs have been refitted to more accurate quantum-mech. energy surfaces. Side-chain torsion parameters have been optimized by fitting to backbone-dependent quantum-mech. energy surfaces, followed by addnl. empirical optimization targeting NMR scalar couplings for unfolded proteins. A comprehensive validation of the revised force field was then performed against a collection of exptl. data: (i) comparison of simulations of eight proteins in their crystal environments with crystal structures; (ii) comparison with backbone scalar couplings for weakly structured peptides; (iii) comparison with NMR residual dipolar couplings and scalar couplings for both backbone and side-chains in folded proteins; (iv) equil. folding of mini-proteins. The results indicate that the revised CHARMM 36 parameters represent an improved model for modeling and simulation studies of proteins, including studies of protein folding, assembly, and functionally relevant conformational changes.
- 41Phillips, J. C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R. D.; Kale, L.; Schulten, K. Scalable molecular dynamics with NAMD J. Comput. Chem. 2005, 26, 1781– 1802 DOI: 10.1002/jcc.20289Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXht1SlsbbJ&md5=189051128443b547f4300a1b8fb0e034Scalable molecular dynamics with NAMDPhillips, James C.; Braun, Rosemary; Wang, Wei; Gumbart, James; Tajkhorshid, Emad; Villa, Elizabeth; Chipot, Christophe; Skeel, Robert D.; Kale, Laxmikant; Schulten, KlausJournal of Computational Chemistry (2005), 26 (16), 1781-1802CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)NAMD is a parallel mol. dynamics code designed for high-performance simulation of large biomol. systems. NAMD scales to hundreds of processors on high-end parallel platforms, as well as tens of processors on low-cost commodity clusters, and also runs on individual desktop and laptop computers. NAMD works with AMBER and CHARMM potential functions, parameters, and file formats. This article, directed to novices as well as experts, first introduces concepts and methods used in the NAMD program, describing the classical mol. dynamics force field, equations of motion, and integration methods along with the efficient electrostatics evaluation algorithms employed and temp. and pressure controls used. Features for steering the simulation across barriers and for calcg. both alchem. and conformational free energy differences are presented. The motivations for and a roadmap to the internal design of NAMD, implemented in C++ and based on Charm++ parallel objects, are outlined. The factors affecting the serial and parallel performance of a simulation are discussed. Finally, typical NAMD use is illustrated with representative applications to a small, a medium, and a large biomol. system, highlighting particular features of NAMD, for example, the Tcl scripting language. The article also provides a list of the key features of NAMD and discusses the benefits of combining NAMD with the mol. graphics/sequence anal. software VMD and the grid computing/collab. software BioCoRE. NAMD is distributed free of charge with source code at www.ks.uiuc.edu.
- 42MacKerell, A. D.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E.; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M. All-atom empirical potential for molecular modeling and dynamics studies of proteins J. Phys. Chem. B 1998, 102, 3586– 3616 DOI: 10.1021/jp973084fGoogle Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 43Feller, S. E.; Zhang, Y.; Pastor, R. W.; Brooks, B. R. Constant pressure molecular dynamics simulation: The Langevin piston method J. Chem. Phys. 1995, 103, 4613– 4621 DOI: 10.1063/1.470648Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXotVentLo%253D&md5=219a4e0a48397a35fa2c62cf99bf225aConstant pressure molecular dynamics simulation: the Langevin piston methodFeller, Scott E.; Zhang, Yuhong; Pastor, Richard W.; Brooks, Bernard R.Journal of Chemical Physics (1995), 103 (11), 4613-21CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A new method for performing mol. dynamics simulations under const. pressure is presented. In the method, which is based on the extended system formalism introduced by Andersen, the deterministic equations of motion for the piston degree of freedom are replaced by a Langevin equation; a suitable choice of collision frequency then eliminates the unphys. "ringing" of the vol. assocd. with the piston mass. In this way it is similar to the "weak coupling algorithm" developed by Berendsen and co-workers to perform mol. dynamics simulation without piston mass effects. It is shown, however, that the weak coupling algorithm induces artifacts into the simulation which can be quite severe for inhomogeneous systems such as aq. biopolymers or liq./liq. interfaces.
- 44Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N·log(N) method for Ewald sums in large systems J. Chem. Phys. 1993, 98, 10089– 10092 DOI: 10.1063/1.464397Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXks1Ohsr0%253D&md5=3c9f230bd01b7b714fd096d4d2e755f6Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 45Andersen, H. C. Rattle: A “velocity” version of the shake algorithm for molecular dynamics calculations J. Comput. Phys. 1983, 52, 24– 34 DOI: 10.1016/0021-9991(83)90014-1Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXjvFOntw%253D%253D&md5=770dfdc612edc5847839ca28ea3d6501RATTLE: a "velocity" version of the SHAKE algorithm for molecular dynamics calculationsAndersen, Hans C.Journal of Computational Physics (1983), 52 (1), 24-34CODEN: JCTPAH; ISSN:0021-9991.An algorithm, called RATTLE, for integrating the equations of motion in mol. dynamics calcns. for mol. models with internal constraints is presented. RATTLE calcs. the positions and velocities at the next time from the positions and velocities at the present time step, without requiring information about the earlier history. It is based on the Verlet algorithm and retains the simplicity of using Cartesian coordinates for each of the atoms to describe the configuration of a mol. with internal constraints. RATTLE guarantees that the coordinates and velocities of the atoms in a mol. satisfy the internal constraints at each time step.
- 46Miyamoto, S.; Kollman, P. A. Settle: An analytical version of the SHAKE and RATTLE algorithm for rigid water models J. Comput. Phys. 1992, 13, 952– 962 DOI: 10.1002/jcc.540130805Google ScholarThere is no corresponding record for this reference.
- 47Hopkins, C. W.; Le Grand, S.; Walker, R. C.; Roitberg, A. E. Long-Time-Step Molecular Dynamics through Hydrogen Mass Repartitioning J. Chem. Theory Comput. 2015, 11, 1864– 1874 DOI: 10.1021/ct5010406Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXksFSrsL0%253D&md5=6ca75b544b63523481de7836d11a3c1bLong-Time-Step Molecular Dynamics through Hydrogen Mass RepartitioningHopkins, Chad W.; Le Grand, Scott; Walker, Ross C.; Roitberg, Adrian E.Journal of Chemical Theory and Computation (2015), 11 (4), 1864-1874CODEN: JCTCCE; ISSN:1549-9618. (American Chemical Society)Previous studies have shown that the method of hydrogen mass repartitioning (HMR) is a potentially useful tool for accelerating mol. dynamics (MD) simulations. By repartitioning the mass of heavy atoms into the bonded hydrogen atoms, it is possible to slow the highest-frequency motions of the macromol. under study, thus allowing the time step of the simulation to be increased by up to a factor of 2. In this communication, we investigate further how this mass repartitioning allows the simulation time step to be increased in a stable fashion without significantly increasing discretization error. To this end, we ran a set of simulations with different time steps and mass distributions on a three-residue peptide to get a comprehensive view of the effect of mass repartitioning and time step increase on a system whose accessible phase space is fully explored in a relatively short amt. of time. We next studied a 129-residue protein, hen egg white lysozyme (HEWL), to verify that the obsd. behavior extends to a larger, more-realistic, system. Results for the protein include structural comparisons from MD trajectories, as well as comparisons of pKa calcns. via const.-pH MD. We also calcd. a potential of mean force (PMF) of a dihedral rotation for the MTS [(1-oxyl-2,2,5,5-tetramethyl-pyrroline-3-methyl)methanethiosulfonate] spin label via umbrella sampling with a set of regular MD trajectories, as well as a set of mass-repartitioned trajectories with a time step of 4 fs. Since no significant difference in kinetics or thermodn. is obsd. by the use of fast HMR trajectories, further evidence is provided that long-time-step HMR MD simulations are a viable tool for accelerating MD simulations for mols. of biochem. interest.
- 48Fiorin, G.; Klein, M. L.; Hénin, J. Using collective variables to drive molecular dynamics simulations Mol. Phys. 2013, 111, 3345– 3362 DOI: 10.1080/00268976.2013.813594Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXpslGit7s%253D&md5=6c1886ae6a2804383260577562fb503bUsing collective variables to drive molecular dynamics simulationsFiorin, Giacomo; Klein, Michael L.; Henin, JeromeMolecular Physics (2013), 111 (22-23), 3345-3362CODEN: MOPHAM; ISSN:0026-8976. (Taylor & Francis Ltd.)A software framework is introduced that facilitates the application of biasing algorithms to collective variables of the type commonly employed to drive massively parallel mol. dynamics (MD) simulations. The modular framework that is presented enables one to combine existing collective variables into new ones, and combine any chosen collective variable with available biasing methods. The latter include the classic time-dependent biases referred to as steered MD and targeted MD, the temp.-accelerated MD algorithm, as well as the adaptive free-energy biases called metadynamics and adaptive biasing force. The present modular software is extensible, and portable between commonly used MD simulation engines.
- 49Comer, J.; Gumbart, J. C.; Hénin, J.; Lelièvre, T.; Pohorille, A.; Chipot, C. The Adaptive Biasing Force Method: Everything You Always Wanted To Know but Were Afraid To Ask J. Phys. Chem. B 2015, 119, 1129– 1151 DOI: 10.1021/jp506633nGoogle Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhsF2nurvP&md5=2ea7fa53559bf7355d24630fc126fcc9The Adaptive Biasing Force Method: Everything You Always Wanted To Know but Were Afraid To AskComer, Jeffrey; Gumbart, James C.; Henin, Jerome; Lelievre, Tony; Pohorille, Andrew; Chipot, ChristopheJournal of Physical Chemistry B (2015), 119 (3), 1129-1151CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)In the host of numerical schemes devised to calc. free energy differences by way of geometric transformations, the adaptive biasing force algorithm has emerged as a promising route to map complex free-energy landscapes. It relies upon the simple concept that as a simulation progresses, a continuously updated biasing force is added to the equations of motion, such that in the long-time limit it yields a Hamiltonian devoid of an av. force acting along the transition coordinate of interest. This means that sampling proceeds uniformly on a flat free-energy surface, thus providing reliable free-energy ests. Much of the appeal of the algorithm to the practitioner is in its phys. intuitive underlying ideas and the absence of any requirements for prior knowledge about free-energy landscapes. Since its inception in 2001, the adaptive biasing force scheme has been the subject of considerable attention, from in-depth math. anal. of convergence properties to novel developments and extensions. The method has also been successfully applied to many challenging problems in chem. and biol. In this contribution, the method is presented in a comprehensive, self-contained fashion, discussing with a crit. eye its properties, applicability, and inherent limitations, as well as introducing novel extensions. Through free-energy calcns. of prototypical mol. systems, many methodol. aspects are examd., from stratification strategies to overcoming the so-called hidden barriers in orthogonal space, relevant not only to the adaptive biasing force algorithm but also to other importance-sampling schemes. On the basis of the discussions in this paper, a no. of good practices for improving the efficiency and reliability of the computed free-energy differences are proposed.
- 50Darve, E.; Rodriguez-Gomez, D.; Pohorille, A. Adaptive biasing force method for scalar and vector free energy calculations J. Chem. Phys. 2008, 128144120 DOI: 10.1063/1.2829861Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXkvFyiu74%253D&md5=6f6eb47d685e873d1ff35ffdc9ae66cbAdaptive biasing force method for scalar and vector free energy calculationsDarve, Eric; Rodriguez-Gomez, David; Pohorille, AndrewJournal of Chemical Physics (2008), 128 (14), 144120/1-144120/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)In free energy calcns. based on thermodn. integration, it is necessary to compute the derivs. of the free energy as a function of one (scalar case) or several (vector case) order parameters. We derive in a compact way a general formulation for evaluating these derivs. as the av. of a mean force acting on the order parameters, which involves first derivs. with respect to both Cartesian coordinates and time. This is in contrast with the previously derived formulas, which require first and second derivs. of the order parameter with respect to Cartesian coordinates. As illustrated in a concrete example, the main advantage of this new formulation is the simplicity of its use, esp. for complicated order parameters. It is also straightforward to implement in a mol. dynamics code, as can be seen from the pseudo-code given at the end. We further discuss how the approach based on time derivs. can be combined with the adaptive biasing force method, an enhanced sampling technique that rapidly yields uniform sampling of the order parameters, and by doing so greatly improves the efficiency of free energy calcns. Using the backbone dihedral angles Φ and Ψ in N-acetylalanyl-N'-methylamide as a numerical example, we present a technique to reconstruct the free energy from its derivs., a calcn. that presents some difficulties in the vector case because of the statistical errors affecting the derivs. (c) 2008 American Institute of Physics.
- 51Hénin, J.; Chipot, C. Overcoming free energy barriers using unconstrained molecular dynamics simulations J. Chem. Phys. 2004, 121, 2904– 2914 DOI: 10.1063/1.1773132Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXmt1Smurk%253D&md5=1f5935c22099d8a213df1249c8ace7bfOvercoming free energy barriers using unconstrained molecular dynamics simulationsHenin, Jerome; Chipot, ChristopheJournal of Chemical Physics (2004), 121 (7), 2904-2914CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)Assocn. of unconstrained mol. dynamics (MD) and the formalisms of thermodn. integration and av. force [Darve and Pohorille, J. Chem. Phys. 115, 9169 (2001)] have been employed to det. potentials of mean force. When implemented in a general MD code, the addnl. computational effort, compared to other std., unconstrained simulations, is marginal. The force acting along a chosen reaction coordinate ξ is estd. from the individual forces exerted on the chem. system and accumulated as the simulation progresses. The estd. free energy deriv. computed for small intervals of ξ is canceled by an adaptive bias to overcome the barriers of the free energy landscape. Evolution of the system along the reaction coordinate is, thus, limited by its sole self-diffusion properties. The illustrative examples of the reversible unfolding of deca-L-alanine, the assocn. of acetate and guanidinium ions in water, the dimerization of methane in water, and its transfer across the water liq.-vapor interface are examd. to probe the efficiency of the method.
Supporting Information
Supporting Information
ARTICLE SECTIONSThe Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b00675.
Effect of type I porcine collagen, up to 2.5 μM, on the absorption spectra of 10 μM RB; representative circular dichroism spectra for a 1.25 μM rose Bengal solution with or without 5.0 μM type I collagen; isothermal titration calorimetry measurements for binding of rose Bengal to a collagen-like peptide (CLP); ζ potential measurements for 2.5 μM type I collagen in 10 mM pH 5.0 MES buffer in the presence of different rose Bengal concentrations; schematic representation for the preparation of RB containing collagen hydrogels; photodecomposition of rose Bengal (125 μM) under oxygen, air, and nitrogen saturated solutions; photodecomposition of RB (125 μM) in the presence of 1.25 mM sodium ascorbate in oxygen, air, and nitrogen saturated solutions; triplet transient lifetime and quenching of RB excited state in the presence of different quenchers; tryptophan degradation mediated by green light exposure in the RB containing collagen matrix (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.