Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Low-Barrier Hydrogen Bond in Fujikurin A–D: A Computational Study
My Activity
  • Open Access
Article

Low-Barrier Hydrogen Bond in Fujikurin A–D: A Computational Study
Click to copy article linkArticle link copied!

  • Hikaru Tanaka
    Hikaru Tanaka
    Department of Chemistry and Biomolecular Science, Faculty of Engineering, Gifu University, 1-1 Yanagido, Gifu 501-1193, Japan
  • Kazuaki Kuwahata
    Kazuaki Kuwahata
    Graduate School of NanobioScience, Yokohama City University, 22-2 Seto, Kanazawa-ku, Yokohama 236-0027, Japan
  • Masanori Tachikawa*
    Masanori Tachikawa
    Graduate School of NanobioScience, Yokohama City University, 22-2 Seto, Kanazawa-ku, Yokohama 236-0027, Japan
    *Email: [email protected]
  • Taro Udagawa*
    Taro Udagawa
    Department of Chemistry and Biomolecular Science, Faculty of Engineering, Gifu University, 1-1 Yanagido, Gifu 501-1193, Japan
    *Email: [email protected]. Phone: +81-58-293-2575. Fax: +81-58-293-2794.
    More by Taro Udagawa
Open PDF

ACS Omega

Cite this: ACS Omega 2022, 7, 16, 14244–14251
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.2c00857
Published April 15, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

The compounds Fujikurin A, B, and D, recently isolated from Fusarium fujikuroi, possess intramolecular low-barrier hydrogen bonds (LBHBs), which are hydrogen bonds with a very low-energy barrier for proton transfer. The isolated compounds have a hydrogen-bonded proton that appears to rapidly switch between two equilibrium states via a transition state (TS). To understand the characteristics of these intramolecular LBHBs in detail, we performed path integral molecular dynamics (PIMD) simulations, which can consider nuclear quantum effects (NQEs) under a finite temperature. The PIMD simulations predicted that the NQE completely washed out the energy barrier for the proton transfer reaction. Consequently, a single-well shape emerged in the results, along with the effective free-energy potential surface for the hydrogen-bonded proton distribution. Thus, we conclude that the hydrogen-bonded proton in Fujikurin does not in fact transfer between two equilibrium structures but widely delocalizes around the global minimum structure involving the TS region.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2022 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Fungi of the genus Fusarium, including Fusarium fujikuroi, produce harmful substances that include mycotoxins. They also induce crop damage with gibberellins that cause the rice disease bakanae. Fusarium fujikuroi is known to produce various secondary metabolites in addition to gibberellins. (1,2) These metabolites have been studied extensively over time. (1−8) Three new compounds, Fujikurin A, B, and D (Figure 1), were recently identified by Bargen et al. These compounds are a secondary metabolite produced by polyketide synthases (PKSs). (6)

Figure 1

Figure 1. Chemical structures of Fujikurin A–D.

Bargen et al. focused on the main product, Fujikurin D, analyzing its structure experimentally (mass spectrometry and NMR) and theoretically (density functional theory, DFT), and determined that it had an intramolecular hydrogen bond. Furthermore, they proposed that the hydrogen-bonded proton may rapidly shift back and forth between two equilibrium structures, EQ(1) and EQ(2) (Figure 2). Moreover, Fujikurin A and B were also found to have intramolecular hydrogen bonds that have the same characteristics as those of Fujikurin D.

Figure 2

Figure 2. Equilibrium and TS structures of Fujikurin D.

However, most of the widely used theoretical calculation methods based on the Born–Oppenheimer approximation do not consider nuclear motion (nuclear quantum effects (NQEs)). The fixed nuclear approach based on the Born–Oppenheimer approximation works well in most situations due to the fact that nuclei are considerably heavier than electrons. However, to analyze systems containing strong hydrogen bonds, including low-barrier hydrogen bonds (LBHBs), where the energy barrier for proton transfer is very low, NQEs are highly useful, enabling the qualitative reproduction of structures. (9−12) Strongly hydrogen-bonded systems have been studied with many helpful methods, including those that involve NQEs. For example, Litman et al. studied porphycene, which has strong intramolecular hydrogen bonds, and reported the temperature dependence of the double hydrogen transfer dynamics, which can be well-understood only if NQEs are taken into account. (13) Ogata et al. investigated asymmetric LBHBs in protonated lysine (LysH+), analyzing the NQEs using the path integral molecular dynamics (PIMD) method. (11) They determined that the hydrogen-bonded proton is not localized on the proton donor or acceptor heavy atoms but is instead delocalized between the heavy atoms from the PIMD simulation.
Thus, the consideration of NQEs is important to correctly understand LBHBs. However, Fujikurin D was analyzed via conventional DFT methods in Bargen’s work, which means that NQEs were not included. Many theoretical methods have been proposed to include NQEs. For example, we have developed multicomponent molecular orbital (MC_MO) methods (9,10,14−16) that go beyond the Born–Oppenheimer approximation to extend the concept of molecular orbitals for electrons to light nuclei. We used MC_MO methods to successfully evaluate the effects of NQEs and H/D isotopes in different hydrogen-bonded systems and proton transfer reactions. PIMD can also consider NQEs statistically within the Born–Oppenheimer approximation framework. (11−13,17−24) The PIMD approach considers NQEs as well as thermal effects. Many PIMD studies have been used to understand NQEs in the context of hydrogen-bonded systems, including LBHBs. In this study, we concentrated on the distribution of hydrogen-bonded protons within Fujikurin A–D. Hence, we carried out PIMD simulations on Fujikurin A–D. We also incorporated typical DFT calculations and conventional MD simulations to elucidate the NQEs and thermal effects in Fujikurin A–D.

Computational Methods

Click to copy section linkSection link copied!

Nuclei are treated as quantum particles in PIMD simulations. In this study, each nucleus was expanded using 16 beads. PIMD simulations were then performed for 90 000 steps, after a thermal equilibration of 10 000 steps using a time step size of 0.1 fs. Furthermore, classical MD (CLMD) simulations, in which each nucleus was represented by one bead, were carried out for 900 000 steps following a thermal equilibration of 100 000 steps using the same time step size for comparison. We performed both CLMD and PIMD simulations in the canonical ensemble (NVT) using a Nosé–Hoover chain (31) at 300 K.
The position of the hydrogen-bonded proton was clarified by defining the proton transfer coordinate (δOH) as
where RO1H and RHO2 are the O1–H covalent bond length and H···O2 hydrogen bond distance, respectively. The notation is illustrated in Figure 3. The value of δOH = 0 implies that the proton is at the center of the oxygen atoms, while a positive or negative δOH value means that the proton is located at a position closer to one of the oxygen atoms.

Figure 3

Figure 3. Hydrogen bond structure of Fujikurin D and definition of atomic labels and coordinates.

The relative effective free energy (ΔF) was calculated from the obtained distributions using the following equation:
where kB, T, and POH) are the Boltzmann constant, temperature, and probability distribution, respectively, as a function of δOH. POHmax)) is the largest distribution of POH) in the histogram, and the free energy in this region is set to 0.0 kcal/mol.

Results and Discussion

Click to copy section linkSection link copied!

Level of Electronic Structure Calculation

The level of electronic structure calculations for the PIMD simulations was based on the relative energy of each stationary point structure of the main product, Fujikurin D. Table 1 shows the relative energies obtained by DFT (CAM-B3LYP), (25) MP2, and semiempirical PM series methods. (26−30)
Table 1. Relative Energies (kcal/mol) of Stationary Point Structures of Intramolecular Proton Transfer Reaction in Fujikurin D Obtained by Several Methods
methodEQ(1)TSEQ(2)
CAM-B3LYP/6-311+G**0.01.81.1
MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G**0.01.90.9
PM60.0  
PM6-DH20.0  
PM6-DH+0.03.22.0
PM6-D3H40.03.01.9
PM70.02.72.5
The MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G** energy was considered to be a highly accurate reference. The CAM-B3LYP/6-311+G** energies were similar to the MP2/6-311+G** energies. Hence, the CAM-B3LYP/6-311+G** method describes the potential energy surface (PES) for the proton-transfer reaction in Fujikurin D accurately. It was also determined that EQ(1) has a lower energy than that of EQ(2). Therefore, EQ(1) is the global minimum structure, and EQ(2) is the metastable equilibrium structure. PIMD simulations require very large numbers of atomic force calculations; therefore, selecting a method that can reproduce highly accurate results with a low computational cost is very important. We focused on a semiempirical PM series method. In Table 1, only the equilibrium state EQ(1) exists on the PM6 and PM6-DH2 PESs, while the proton-transfer equilibrium state EQ(2) and the TS exist on the PM6-DH+, PM6-D3H4, and PM7 PESs. In particular, the PM6-D3H4 method reproduced the relative energies obtained by the MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G** method.
Next, we confirmed the validity of the PM6-D3H4 method for Fujikurins A and B. The relative energies of Fujikurin A and B are listed in Table 2.
Table 2. Relative Energies (kcal/mol) of Stationary Point Structures of Intramolecular Proton-Transfer Reaction in Fujikurin A and B Obtained by Several Methods
 Fujikurin AFujikurin B
methodEQ(1)TSEQ(2)EQ(1)TSEQ(2)
CAM-B3LYP/6-311+G**0.01.40.20.02.31.7
MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G**0.01.4–0.20.02.21.4
PM6-D3H40.02.81.70.03.72.9
Similar to the case of Fujikurin D, the CAM-B3LYP/6-311+G** relative energies are like those calculated by MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G**, indicating that the CAM-B3LYP-optimized geometries were adequate for Fujikurin A and B. The MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G** activation barriers of Fujikurin A and Fujikurin B were 0.5 kcal/mol lower and 0.3 kcal/mol higher than that of Fujikurin D, respectively. Moreover, the PM6-D3H4 method estimated the activation barriers of Fujikurin A and Fujikurin B as 0.2 kcal/mol lower and 0.7 kcal/mol higher than those of Fujikurin D, respectively. This implies that the PM6-D3H4 method is also adequate for analyzing the intramolecular proton-transfer reaction in Fujikurin A and B. Finally, we used the PM6-D3H4 method to calculate atomic forces.

Structural Features of Fujikurin D

For Fujikurin D, Figure 4 shows the one-dimensional distributions of RO1H, RHO2, and RO1O2, while Table 3 has the average values of RO1H, RHO2, and RO1O2 obtained by CLMD and PIMD simulations. The values shown in parentheses are the statistical errors of the average bond lengths calculated using the block average method. (32)Table 3 lists the optimized RO1H, RHO2, and RO1O2 values in the equilibrium structure.

Figure 4

Figure 4. One-dimensional distributions of (a) RO1H, (b) RHO2, and (c) RO1O2 obtained by CLMD and PIMD simulations. The dashed lines represent the values of two equilibrium structures.

Table 3. Average Values of RO1H (Å), RHO2 (Å), and RO1O2 (Å) Obtained by CLMD and PIMD Simulations and the Interatomic Distances in Equilibrium Structures EQ(1) and EQ(2)
 RO1HRHO2RO1O2
CLMD1.103(7)1.607(8)2.525(2)
PIMD1.112(4)1.601(10)2.512(6)
EQ(1)1.0911.5822.504
EQ(2)1.4201.1742.440
In a comparison of the CLMD and PIMD results for the one-dimensional distributions of RO1H and RHO2 (Figure 4a,b), the distributions generated by PIMD simulations are much wider than those produced by CLMD simulations owing to quantum fluctuations. The results of the CLMD simulations showed two peaks. Because EQ(1) is more stable than EQ(2), the peak around EQ(1) is much larger. However, the distributions for the CLMD and PIMD simulations for RO1O2 are similar. Thus, the impact of the NQEs on RO1O2 was smaller than that on RO1H.
We focused on the covalent RO1H average bond lengths in Table 3 and found that the relationship of the RO1H values is as follows:
The average value obtained in the CLMD simulation was greater than the ROH in EQ(1). The CLMD simulation includes thermal effects only; therefore, the elongation of the covalent RO1H bond length is due to thermal motion. The PIMD simulation, which includes both thermal effects and NQEs, provided a greater RO1H average bond length than that of the CLMD simulation. Thus, the NQEs stretch the covalent bond length. These results are often observed in other hydrogen-bonded systems. (11,12)
However, the average values of RHO2 and RO1O2 in Table 3 show the following relationship, which is different from that for the covalent ROH lengths:
The bond lengths obtained from the CLMD and PIMD simulations were greater than those in EQ(1) because of the thermal effect, as in the case of RO1H. To explain why the RHO2 and RO1O2 values obtained in the PIMD simulations were smaller than those obtained in the CLMD simulation, we focused on the Mulliken charges of oxygen and hydrogen atoms (Figure 5).

Figure 5

Figure 5. Mulliken charges of the oxygen and the hydrogen atoms obtained in (a) CLMD and (b) PIMD simulations.

The Mulliken charge of H was found to be +0.426 in the CLMD result, whereas it was +0.430 from PIMD, which means that the hydrogen-bonded proton in the PIMD simulation was more positively charged than that in the CLMD simulation. The finite spread of the nuclei in the PIMD simulation induced the NQEs, which affected the surrounding electron distributions. We can find that the proton acceptor, O2, is more negative than O1. In addition, O2 in the PIMD simulation is less negative than that in the CLMD one, unlike O1. Since the proton donor and acceptor atoms swap their roles by the proton transfer, O1 and O2 become more and less negative around the TS region. The hydrogen-bonded proton can access the TS region more easily in the PIMD simulation rather than in the CLMD one, as described below. Thus, O2 seems to be less negative in the PIMD simulation than in the CLMD one. We should focus on the sum of the negative charges of O1 and O2 for the strength of the interaction. The greater negative charge is found in the PIMD simulation than in the CLMD. The more positively charged hydrogen-bonded proton interacts more strongly with the negatively charged oxygen atoms in the hydrogen bond moiety. Thus, the distance between the oxygen atoms in the PIMD result was smaller than that in the CLMD result owing to the strong hydrogen bond interaction.

Relative Free-Energy Potential Surface for Proton-Transfer Coordinate

To understand the hydrogen-bonded proton distribution, we focused on the one-dimensional relative free-energy potential surface for the proton-transfer coordinate δOH. Figure 6 shows the one-dimensional relative effective free-energy potential surfaces from the CLMD and PIMD simulations.

Figure 6

Figure 6. One-dimensional relative effective free-energy landscapes of δOH obtained by CLMD and PIMD simulations. The dashed lines represent two equilibrium structures.

First, we focused on the CLMD results. The relative effective free-energy potential obtained by the CLMD simulation is a double-well potential with two minima. The positions of the minima are in good agreement with the δOH values of the equilibrium structures shown by the black dashed lines in Figure 6. Thus, it is difficult to overcome the activation energy barrier using thermal effects alone. The barrier height of the free-energy potential in the CLMD result was estimated as 3.6 kcal/mol, which is slightly larger than that from the static electronic structure calculation, 3.0 kcal/mol (Table 1). This trend has also been reported for other LBHB systems. (11,17) However, the energy barrier completely disappeared in the PIMD results, and the shape of the free-energy potential was a single well. The PIMD simulation took the thermal effect and NQEs into account; thus, the energy barrier of the proton transfer was washed out by the NQEs. Similar NQE influences on the proton distribution around the TS structure were reported for LBHB in LysH+ by Ogata et al. (11) In particular, the NQEs on proton transfer might be the primary factor because the NQEs of light nuclei were more prominent than those of heavier nuclei. It should be noted here that, in a symmetrical hydrogen-bonded system such as a maleate anion, (12) the potential minimum should be at δOH = 0 Å. However, in Fujikurin D, the minimum potential is located at approximately δOH = −0.5 Å, which corresponds to the global minimum structure EQ(1), because Fujikurin D has an asymmetric PES in the proton-transfer coordinate.
The two-dimensional free-energy landscapes of δOH and ROO are shown in Figure 7 to aid a better understanding of the proton distribution. There are two minima located around the equilibrium structures in the free-energy landscape simulated using the CLMD method (Figure 7a). Thus, the equilibrium structure analogues, in the which transferring proton is localized to either oxygen atom, are frequently observed as expected from the one-dimensional free-energy landscape (Figure 6). In addition, the energy maximum exists near the TS structure. Therefore, the distribution near the TS region is sparse compared to those around the equilibrium structures. Moreover, the low-energy regions are distributed along the IRC pathways. Thus, the hydrogen-bonded proton is mainly distributed around the equilibrium structures, and proton transfer may occur between the two equilibrium structures. Thus, the predictions of the CLMD are consistent with those by Bargen.

Figure 7

Figure 7. Two-dimensional free-energy landscapes of δOH and ROO obtained with (a) CLMD and (b) PIMD. Triangles and circles represent the equilibrium and the TS structures, respectively, and dashed lines are IRC pathways. Free energy is in units of kcal/mol.

In contrast, the PIMD simulation results are qualitatively different from the CLMD results. As seen in Figure 7b, the energy barrier for the proton transfer was completely overcome by the NQEs. The PIMD results show a broad single-well potential centered at the global minimum (δOH = −0.5 Å) and the corresponding one-dimensional free-energy potential (Figure 6). The regions around the TS and metastable structures are also included in this broad single-well potential. Unlike in the CLMD predictions, the proton distribution extends to regions far from the IRC. This corner-cutting phenomenon is induced by the NQEs and is seen in several systems, such as N2H7+. (18) Our PIMD simulation shows a broad single-well potential centered on the global minimum structure for Fujikurin D. In other words, in the PIMD predictions, the proton overcomes the proton-transfer barrier freely and is not localized around either oxygen atoms, in contrast to Bargen’s prediction of fast transfer between the two equilibrium structures.

Proton Coordinate

We defined a coordinate system to analyze how the hydrogen-bonded protons are distributed, as shown in Figure 8. The origin is the center of mass of two oxygen atoms, and the xy-plane is defined to be in the O1–O2–C plane. Proton-distribution landscapes were obtained in three dimensions from both CLMD (Figure 8a) and PIMD (Figure 8b). Simulations in this coordinate system are shown in Figure 8.

Figure 8

Figure 8. Three-dimensional distribution of hydrogen-bonded proton in Fujikurin D. The origin is defined as the center of mass of the O atoms and the xy-plane as the O1–O2–C plane.

The CLMD results are shown in Figure 8a. Many structures are located around the two equilibrium structures, which were indicated by two peaks exhibited by the proton distribution in the xy- and xz-planes. In contrast, the PIMD result (Figure 8b) shows single broad distributions in every direction, depicting the activation barrier wash out caused by the NQEs, as mentioned above. While there are pronounced differences in the x- and y-directions when PIMD is compared to CLMD, the changes in the z-axis direction are not significant. To illustrate the changes more clearly, we show the difference distribution landscapes, which are the differences between the proton distributions obtained in the CLMD and PIMD simulations, in Figure 9.

Figure 9

Figure 9. Difference distribution landscapes of proton distributions obtained from CLMD and PIMD simulations of Fujikurin D. Red (blue) indicates regions where the distribution in the PIMD (CLMD) simulated results is denser than that in the CLMD (PIMD) simulated results.

In Figure 9, red indicates regions where the distributions in the PIMD results are denser than those in the CLMD results. This red area is spread in the xy-plane. Thus, distributions in the x- and y-directions increase because of the NQEs. However, in the yz- and zx-planes, the distribution in the z-direction did not increase much compared to the x- and y-directions. This means the NQEs did not considerably affect the protonic motion along the z-axis direction. The presence of the intramolecular hydrogen bond is considered to be a reason for the directional behavior of the NQE. The NQEs influence the electronic distribution, and as a result, the hydrogen bond interaction was enhanced in the PIMD simulations. Proton donor and acceptor atoms are always in the xy-plane in this environment. Hence, the hydrogen-bonded proton was attracted more strongly from the x- and y-directions than from the z-direction. This explains why the NQEs are pronounced in only the x- and y-directions.

Proton Distribution in Fujikurin A and B

Figure 10a,b shows the two-dimensional free-energy values, δOH and ROO, for Fujikurin A and B, respectively, to demonstrate how the distribution of the transferring proton is affected by the height of the activation barrier.

Figure 10

Figure 10. Two-dimensional free-energy landscapes δOH and ROO of Fujikurin A and B for (a) CLMD and (b) PIMD. Triangles and circles represent equilibrium and the TS structures, respectively, and dashed lines represent IRC pathways. Free energy is expressed in units of kcal/mol.

The CLMD results exhibit two minima, while the PIMD results show one minimum. This means the NQEs washed out the activation barrier in all Fujikurin compounds. Among the Fujikurin A–D compounds, Fujikurin A has the lowest activation barrier and largest low-energy region (represented by orange and yellow) in the two-dimensional free-energy results obtained by PIMD. NQEs extend the low-energy region around the TS, as well as near the metastable equilibrium structure.
It is worth noting that the PM6-D3H4 calculations overestimate the activation barrier slightly, when compared to those of the MP2/aug-cc-pVDZ//CAM-B3LYP/6-311+G** method. Hence, the distributions near the TS and metastable equilibrium structure may be denser on the more precise potential energy hypersurface.

Conclusion

Click to copy section linkSection link copied!

In this study, we investigated the NQE and thermal effects on Fujikurin A–D intramolecular hydrogen bonds via PIMD simulation, which takes the thermal effect and the NQE into account.
For Fujikurin D, the CLMD simulations demonstrate that the hydrogen-bonded proton is rapidly switching between two equilibrium states, as predicted in a previous study by Bargen et al. In contrast, the PIMD simulation results are qualitatively different because NQEs are taken into consideration. The results imply a delocalized distribution on the single-well potential as a result of the wash out of the proton-transfer energy barrier. Furthermore, the NQE in Fujikurin D was found to be direction-dependent. The PIMD simulations demonstrated that the change in the PES of the proton-transfer reaction in Fujikurin D after considering NQEs could be observed even qualitatively.
Among Fujikurin A, B, and D, the widest proton distribution was found in Fujikurin A, which has the lowest activation barrier for proton transfer. In Fujikurin A, protons are distributed around the TS, as well as in the metastable equilibrium structure. Because the activation barriers were overestimated slightly by the PM6-D3H4 level of calculation, the proton distribution near the TS should actually be denser than calculated in this study. Our results can help elucidate the structures and properties of new secondary metabolites of Fujikurins.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Hikaru Tanaka - Department of Chemistry and Biomolecular Science, Faculty of Engineering, Gifu University, 1-1 Yanagido, Gifu 501-1193, Japan
    • Kazuaki Kuwahata - Graduate School of NanobioScience, Yokohama City University, 22-2 Seto, Kanazawa-ku, Yokohama 236-0027, Japan
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This work was supported by JSPS KAKENHI Grants 21J00806 (to K.K.); 18H01945, 19H05063, and 21H00026 (to M.T.); and 20H05739 and 21K04991 (to T.U.). The generous allotment of computational resources from the Research Center for Computational Science, National Institute of Natural Science, Japan, is also gratefully acknowledged.

References

Click to copy section linkSection link copied!

This article references 32 other publications.

  1. 1
    Yabuta, T.; Kambe, K.; Hayashi, T. Biochemistry of the ″bakanae″ fungus. I. Fusarinic acid, a new product of the “bakanae” fungus. Nippon Nogei Kagaku Kaishi 1934, 10, 10591068
  2. 2
    Graebe, J. E. Gibberellin Biosynthesis and Control. Annu. Rev. Plant. Physiol. 1987, 38, 419,  DOI: 10.1146/annurev.pp.38.060187.002223
  3. 3
    Brady, S. F.; Clardy, J. CR377, a New Pentaketide Antifungal Agent Isolated from an Endophytic Fungus. J. Nat. Prod. 2000, 63, 14471448,  DOI: 10.1021/np990568p
  4. 4
    Niehaus, E. M.; Kleigrewe, K.; Wiemann, P.; Studt, L.; Sieber, C. M. K.; Connolly, L. R.; Freitag, M.; Güldener, U.; Tudzynski, B.; Humpf, H. U. Genetic Manipulation of the Fusarium Fujikuroi Fusarin Gene Cluster Yields Insight into the Complex Regulation and Fusarin Biosynthetic Pathway. Chem. Biol. 2013, 20, 10551066,  DOI: 10.1016/j.chembiol.2013.07.004
  5. 5
    Niehaus, E. M.; von Bargen, K. W.; Espino, J. J.; Pfannmüller, A.; Humpf, H. U.; Tudzynski, B. Characterization of the Fusaric Acid Gene Cluster in Fusarium Fujikuroi. Appl. Microbiol. Biotechnol. 2014, 98, 17491762,  DOI: 10.1007/s00253-013-5453-1
  6. 6
    von Bargen, K. W.; Niehaus, E. M.; Krug, I.; Bergander, K.; Würthwein, E. U.; Tudzynski, B.; Humpf, H. U. Isolation and Structure Elucidation of Fujikurins A-D: Products of the PKS19 Gene Cluster in Fusarium Fujikuroi. J. Nat. Prod. 2015, 78, 18091815,  DOI: 10.1021/np5008137
  7. 7
    Niehaus, E. M.; Kim, H. K.; Münsterkötter, M.; Janevska, S.; Arndt, B.; Kalinina, S. A.; Houterman, P. M.; Ahn, I. P.; Alberti, I.; Tonti, S.; Kim, D. W.; Sieber, C. M. K.; Humpf, H. U.; Yun, S. H.; Güldener, U.; Tudzynski, B. Comparative Genomics of Geographically Distant Fusarium Fujikuroi Isolates Revealed Two Distinct Pathotypes Correlating with Secondary Metabolite Profiles. PLoS Pathogens 2017, 13, e1006670  DOI: 10.1371/journal.ppat.1006670
  8. 8
    Wang, W. G.; Wang, H.; Du, L. Q.; Li, M.; Chen, L.; Yu, J.; Cheng, G. G.; Zhan, M. T.; Hu, Q. F.; Zhang, L.; Yao, M.; Matsuda, Y. Molecular Basis for the Biosynthesis of an Unusual Chain-Fused Polyketide, Gregatin A. J. Am. Chem. Soc. 2020, 142, 84648472,  DOI: 10.1021/jacs.0c02337
  9. 9
    Moreno, D. v.; González, S. A.; Reyes, A. Turning Symmetric an Asymmetric Hydrogen Bond with the Inclusion of NQEs: The Case of the [CN···H···NC] Complex. J. Chem. Phys. 2011, 134, 024115,  DOI: 10.1063/1.3521272
  10. 10
    Kanematsu, Y.; Tachikawa, M. Theoretical Analysis of Geometry and NMR Isotope Shift in Hydrogen-Bonding Center of Photoactive Yellow Protein by Combination of Multicomponent Quantum Mechanics and ONIOM Scheme. J. Chem. Phys. 2014, 141, 185101,  DOI: 10.1063/1.4900987
  11. 11
    Ogata, Y.; Daido, M.; Kawashima, Y.; Tachikawa, M. NQEs on Protonated Lysine with an Asymmetric Low Barrier Hydrogen Bond: An Ab Initio Path Integral Molecular Dynamics Study. RSC Adv. 2013, 3, 2525225257,  DOI: 10.1039/c3ra44077j
  12. 12
    Kawashima, Y.; Tachikawa, M. Ab Initio Path Integral Molecular Dynamics Study of the NQE on Out-of-Plane Ring Deformation of Hydrogen Maleate Anion. J. Chem. Theory Comput. 2014, 10, 153163,  DOI: 10.1021/ct4007986
  13. 13
    Litman, Y.; Richardson, J. O.; Kumagai, T.; Rossi, M. Elucidating the Nuclear Quantum Dynamics of Intramolecular Double Hydrogen Transfer in Porphycene. J. Am. Chem. Soc. 2019, 141, 25262534,  DOI: 10.1021/jacs.8b12471
  14. 14
    Udagawa, T.; Tachikawa, M. H/D Isotope Effect on Porphine and Porphycene Molecules with Multicomponent Hybrid Density Functional Theory. J. Chem. Phys. 2006, 125, 244105,  DOI: 10.1063/1.2403857
  15. 15
    Funahashi, H.; Tachikawa, M.; Udagawa, T. Determining If Reaction Selectivity Can Be Controlled by the H/D Isotope Effect in CH···O Interactions. Org. Lett. 2020, 22, 94399443,  DOI: 10.1021/acs.orglett.0c03351
  16. 16
    Udagawa, T.; Murphy, R. B.; Darwish, T. A.; Tachikawa, M.; Mori, S. H/D Isotope Effects in Keto-Enol Tautomerism of β-Dicarbonyl Compounds ─Importance of NQEs of Hydrogen Nuclei─. Bull. Chem. Soc. Jpn. 2021, 94, 19541962,  DOI: 10.1246/bcsj.20210083
  17. 17
    Tuckerman, M. E.; Marx, D.; Klein, M. L.; Parrinello, M. On the Quantum Nature of the Shared Proton in Hydrogen Bonds. Science 1997, 275, 817820,  DOI: 10.1126/science.275.5301.817
  18. 18
    Ishibashi, H.; Hayashi, A.; Shiga, M.; Tachikawa, M. Geometric Isotope Effect on the N2H7+ Cation and N2H5 Anion by Ab Initio Path Integral Molecular Dynamics Simulation. ChemPhysChem 2008, 9, 383387,  DOI: 10.1002/cphc.200700570
  19. 19
    Durlak, P.; Latajka, Z. Car-Parrinello and Path Integral Molecular Dynamics Study of the Intramolecular Hydrogen Bonds in the Crystals of Benzoylacetone and dideuterobenzoylacetone. Phys. Chem. Chem. Phys. 2014, 16, 2302623037,  DOI: 10.1039/C4CP02569E
  20. 20
    Etinski, M.; Ensing, B. Puzzle of the Intramolecular Hydrogen Bond of Dibenzoylmethane Resolved by Molecular Dynamics Simulations. J. Phys. Chem. A 2018, 122, 59455954,  DOI: 10.1021/acs.jpca.8b01930
  21. 21
    Durlak, P.; Latajka, Z. Investigations of the Hydrogen Bond in the Crystals of Tropolone and Thiotropolone via Car-Parrinello and Path Integral Molecular Dynamics. J. Comput. Chem. 2018, 40, 671687,  DOI: 10.1002/jcc.25753
  22. 22
    Sauceda, H. E.; Vassilev-Galindo, V.; Chmiela, S.; Müller, K. R.; Tkatchenko, A. Dynamical Strengthening of Covalent and Non-Covalent Molecular Interactions by NQEs at Finite Temperature. Nat. Commun. 2021, 12 (1), 442,  DOI: 10.1038/s41467-020-20212-1
  23. 23
    Kuwahata, K.; Tachikawa, M. Path Integral Molecular Dynamics Study on NH 4+(H2O). Few-Body Sys. 2021, 62, 96,  DOI: 10.1007/s00601-021-01689-y
  24. 24
    Udagawa, T.; Kuwahata, K.; Tachikawa, M. Competitive NQE and H/D Isotope Effect on Torsional Motion of H2O2: An Ab Initio Path Integral Molecular Dynamics Study. Comput. Theor. Chem. 2022, 1208, 113542,  DOI: 10.1016/j.comptc.2021.113542
  25. 25
    Yanai, T.; Tew, D. P.; Handy, N. C. A New Hybrid Exchange-Correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 5157,  DOI: 10.1016/j.cplett.2004.06.011
  26. 26
    Stewart, J. J. P. Optimization of Parameters for Semiempirical Methods V: Modification of NDDO Approximations and Application to 70 Elements. J. Mol. Model. 2007, 13, 11731213,  DOI: 10.1007/s00894-007-0233-4
  27. 27
    Korth, M.; Pitoňák, M.; Řezáč, J.; Hobza, P. A Transferable H-Bonding Correction for Semiempirical Quantum-Chemical Methods. J. Chem. Theory Comput. 2010, 6, 344352,  DOI: 10.1021/ct900541n
  28. 28
    Korth, M. Third-Generation Hydrogen-Bonding Corrections for Semiempirical QM Methods and Force Fields. J. Chem. Theroy Comput. 2010, 6, 38083816,  DOI: 10.1021/ct100408b
  29. 29
    Řezáč, J.; Hobza, P. Advanced Corrections of Hydrogen Bonding and Dispersion for Semiempirical Quantum Mechanical Methods. J. Chem. Theory Comput. 2012, 8, 141151,  DOI: 10.1021/ct200751e
  30. 30
    Stewart, J. J. P. Optimization of Parameters for Semiempirical Methods VI: More Modifications to the NDDO Approximations and Re-Optimization of Parameters. J. Mol. Model. 2013, 19, 132,  DOI: 10.1007/s00894-012-1667-x
  31. 31
    Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé-Hoover Chains: The Canonical Ensemble via Continuous Dynamics. J. Chem. Phys. 1992, 97, 26352643,  DOI: 10.1063/1.463940
  32. 32
    Flyvbjerg, H.; Petersen, H. G. Error Estimates on Averages of Correlated Data. J. Chem. Phys. 1989, 91, 461466,  DOI: 10.1063/1.457480

Cited By

Click to copy section linkSection link copied!

This article is cited by 1 publications.

  1. Taro Udagawa, Hikaru Tanaka, Kazuaki Kuwahata, Masanori Tachikawa. Location of the Shared Proton in Proton-Bound Dimer Compound of Hydrogen Sulfate and Formate: Path Integral Molecular Dynamics Study. The Journal of Physical Chemistry A 2024, 128 (11) , 2103-2110. https://doi.org/10.1021/acs.jpca.4c00239
Open PDF

ACS Omega

Cite this: ACS Omega 2022, 7, 16, 14244–14251
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.2c00857
Published April 15, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

821

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Chemical structures of Fujikurin A–D.

    Figure 2

    Figure 2. Equilibrium and TS structures of Fujikurin D.

    Figure 3

    Figure 3. Hydrogen bond structure of Fujikurin D and definition of atomic labels and coordinates.

    Figure 4

    Figure 4. One-dimensional distributions of (a) RO1H, (b) RHO2, and (c) RO1O2 obtained by CLMD and PIMD simulations. The dashed lines represent the values of two equilibrium structures.

    Figure 5

    Figure 5. Mulliken charges of the oxygen and the hydrogen atoms obtained in (a) CLMD and (b) PIMD simulations.

    Figure 6

    Figure 6. One-dimensional relative effective free-energy landscapes of δOH obtained by CLMD and PIMD simulations. The dashed lines represent two equilibrium structures.

    Figure 7

    Figure 7. Two-dimensional free-energy landscapes of δOH and ROO obtained with (a) CLMD and (b) PIMD. Triangles and circles represent the equilibrium and the TS structures, respectively, and dashed lines are IRC pathways. Free energy is in units of kcal/mol.

    Figure 8

    Figure 8. Three-dimensional distribution of hydrogen-bonded proton in Fujikurin D. The origin is defined as the center of mass of the O atoms and the xy-plane as the O1–O2–C plane.

    Figure 9

    Figure 9. Difference distribution landscapes of proton distributions obtained from CLMD and PIMD simulations of Fujikurin D. Red (blue) indicates regions where the distribution in the PIMD (CLMD) simulated results is denser than that in the CLMD (PIMD) simulated results.

    Figure 10

    Figure 10. Two-dimensional free-energy landscapes δOH and ROO of Fujikurin A and B for (a) CLMD and (b) PIMD. Triangles and circles represent equilibrium and the TS structures, respectively, and dashed lines represent IRC pathways. Free energy is expressed in units of kcal/mol.

  • References


    This article references 32 other publications.

    1. 1
      Yabuta, T.; Kambe, K.; Hayashi, T. Biochemistry of the ″bakanae″ fungus. I. Fusarinic acid, a new product of the “bakanae” fungus. Nippon Nogei Kagaku Kaishi 1934, 10, 10591068
    2. 2
      Graebe, J. E. Gibberellin Biosynthesis and Control. Annu. Rev. Plant. Physiol. 1987, 38, 419,  DOI: 10.1146/annurev.pp.38.060187.002223
    3. 3
      Brady, S. F.; Clardy, J. CR377, a New Pentaketide Antifungal Agent Isolated from an Endophytic Fungus. J. Nat. Prod. 2000, 63, 14471448,  DOI: 10.1021/np990568p
    4. 4
      Niehaus, E. M.; Kleigrewe, K.; Wiemann, P.; Studt, L.; Sieber, C. M. K.; Connolly, L. R.; Freitag, M.; Güldener, U.; Tudzynski, B.; Humpf, H. U. Genetic Manipulation of the Fusarium Fujikuroi Fusarin Gene Cluster Yields Insight into the Complex Regulation and Fusarin Biosynthetic Pathway. Chem. Biol. 2013, 20, 10551066,  DOI: 10.1016/j.chembiol.2013.07.004
    5. 5
      Niehaus, E. M.; von Bargen, K. W.; Espino, J. J.; Pfannmüller, A.; Humpf, H. U.; Tudzynski, B. Characterization of the Fusaric Acid Gene Cluster in Fusarium Fujikuroi. Appl. Microbiol. Biotechnol. 2014, 98, 17491762,  DOI: 10.1007/s00253-013-5453-1
    6. 6
      von Bargen, K. W.; Niehaus, E. M.; Krug, I.; Bergander, K.; Würthwein, E. U.; Tudzynski, B.; Humpf, H. U. Isolation and Structure Elucidation of Fujikurins A-D: Products of the PKS19 Gene Cluster in Fusarium Fujikuroi. J. Nat. Prod. 2015, 78, 18091815,  DOI: 10.1021/np5008137
    7. 7
      Niehaus, E. M.; Kim, H. K.; Münsterkötter, M.; Janevska, S.; Arndt, B.; Kalinina, S. A.; Houterman, P. M.; Ahn, I. P.; Alberti, I.; Tonti, S.; Kim, D. W.; Sieber, C. M. K.; Humpf, H. U.; Yun, S. H.; Güldener, U.; Tudzynski, B. Comparative Genomics of Geographically Distant Fusarium Fujikuroi Isolates Revealed Two Distinct Pathotypes Correlating with Secondary Metabolite Profiles. PLoS Pathogens 2017, 13, e1006670  DOI: 10.1371/journal.ppat.1006670
    8. 8
      Wang, W. G.; Wang, H.; Du, L. Q.; Li, M.; Chen, L.; Yu, J.; Cheng, G. G.; Zhan, M. T.; Hu, Q. F.; Zhang, L.; Yao, M.; Matsuda, Y. Molecular Basis for the Biosynthesis of an Unusual Chain-Fused Polyketide, Gregatin A. J. Am. Chem. Soc. 2020, 142, 84648472,  DOI: 10.1021/jacs.0c02337
    9. 9
      Moreno, D. v.; González, S. A.; Reyes, A. Turning Symmetric an Asymmetric Hydrogen Bond with the Inclusion of NQEs: The Case of the [CN···H···NC] Complex. J. Chem. Phys. 2011, 134, 024115,  DOI: 10.1063/1.3521272
    10. 10
      Kanematsu, Y.; Tachikawa, M. Theoretical Analysis of Geometry and NMR Isotope Shift in Hydrogen-Bonding Center of Photoactive Yellow Protein by Combination of Multicomponent Quantum Mechanics and ONIOM Scheme. J. Chem. Phys. 2014, 141, 185101,  DOI: 10.1063/1.4900987
    11. 11
      Ogata, Y.; Daido, M.; Kawashima, Y.; Tachikawa, M. NQEs on Protonated Lysine with an Asymmetric Low Barrier Hydrogen Bond: An Ab Initio Path Integral Molecular Dynamics Study. RSC Adv. 2013, 3, 2525225257,  DOI: 10.1039/c3ra44077j
    12. 12
      Kawashima, Y.; Tachikawa, M. Ab Initio Path Integral Molecular Dynamics Study of the NQE on Out-of-Plane Ring Deformation of Hydrogen Maleate Anion. J. Chem. Theory Comput. 2014, 10, 153163,  DOI: 10.1021/ct4007986
    13. 13
      Litman, Y.; Richardson, J. O.; Kumagai, T.; Rossi, M. Elucidating the Nuclear Quantum Dynamics of Intramolecular Double Hydrogen Transfer in Porphycene. J. Am. Chem. Soc. 2019, 141, 25262534,  DOI: 10.1021/jacs.8b12471
    14. 14
      Udagawa, T.; Tachikawa, M. H/D Isotope Effect on Porphine and Porphycene Molecules with Multicomponent Hybrid Density Functional Theory. J. Chem. Phys. 2006, 125, 244105,  DOI: 10.1063/1.2403857
    15. 15
      Funahashi, H.; Tachikawa, M.; Udagawa, T. Determining If Reaction Selectivity Can Be Controlled by the H/D Isotope Effect in CH···O Interactions. Org. Lett. 2020, 22, 94399443,  DOI: 10.1021/acs.orglett.0c03351
    16. 16
      Udagawa, T.; Murphy, R. B.; Darwish, T. A.; Tachikawa, M.; Mori, S. H/D Isotope Effects in Keto-Enol Tautomerism of β-Dicarbonyl Compounds ─Importance of NQEs of Hydrogen Nuclei─. Bull. Chem. Soc. Jpn. 2021, 94, 19541962,  DOI: 10.1246/bcsj.20210083
    17. 17
      Tuckerman, M. E.; Marx, D.; Klein, M. L.; Parrinello, M. On the Quantum Nature of the Shared Proton in Hydrogen Bonds. Science 1997, 275, 817820,  DOI: 10.1126/science.275.5301.817
    18. 18
      Ishibashi, H.; Hayashi, A.; Shiga, M.; Tachikawa, M. Geometric Isotope Effect on the N2H7+ Cation and N2H5 Anion by Ab Initio Path Integral Molecular Dynamics Simulation. ChemPhysChem 2008, 9, 383387,  DOI: 10.1002/cphc.200700570
    19. 19
      Durlak, P.; Latajka, Z. Car-Parrinello and Path Integral Molecular Dynamics Study of the Intramolecular Hydrogen Bonds in the Crystals of Benzoylacetone and dideuterobenzoylacetone. Phys. Chem. Chem. Phys. 2014, 16, 2302623037,  DOI: 10.1039/C4CP02569E
    20. 20
      Etinski, M.; Ensing, B. Puzzle of the Intramolecular Hydrogen Bond of Dibenzoylmethane Resolved by Molecular Dynamics Simulations. J. Phys. Chem. A 2018, 122, 59455954,  DOI: 10.1021/acs.jpca.8b01930
    21. 21
      Durlak, P.; Latajka, Z. Investigations of the Hydrogen Bond in the Crystals of Tropolone and Thiotropolone via Car-Parrinello and Path Integral Molecular Dynamics. J. Comput. Chem. 2018, 40, 671687,  DOI: 10.1002/jcc.25753
    22. 22
      Sauceda, H. E.; Vassilev-Galindo, V.; Chmiela, S.; Müller, K. R.; Tkatchenko, A. Dynamical Strengthening of Covalent and Non-Covalent Molecular Interactions by NQEs at Finite Temperature. Nat. Commun. 2021, 12 (1), 442,  DOI: 10.1038/s41467-020-20212-1
    23. 23
      Kuwahata, K.; Tachikawa, M. Path Integral Molecular Dynamics Study on NH 4+(H2O). Few-Body Sys. 2021, 62, 96,  DOI: 10.1007/s00601-021-01689-y
    24. 24
      Udagawa, T.; Kuwahata, K.; Tachikawa, M. Competitive NQE and H/D Isotope Effect on Torsional Motion of H2O2: An Ab Initio Path Integral Molecular Dynamics Study. Comput. Theor. Chem. 2022, 1208, 113542,  DOI: 10.1016/j.comptc.2021.113542
    25. 25
      Yanai, T.; Tew, D. P.; Handy, N. C. A New Hybrid Exchange-Correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 5157,  DOI: 10.1016/j.cplett.2004.06.011
    26. 26
      Stewart, J. J. P. Optimization of Parameters for Semiempirical Methods V: Modification of NDDO Approximations and Application to 70 Elements. J. Mol. Model. 2007, 13, 11731213,  DOI: 10.1007/s00894-007-0233-4
    27. 27
      Korth, M.; Pitoňák, M.; Řezáč, J.; Hobza, P. A Transferable H-Bonding Correction for Semiempirical Quantum-Chemical Methods. J. Chem. Theory Comput. 2010, 6, 344352,  DOI: 10.1021/ct900541n
    28. 28
      Korth, M. Third-Generation Hydrogen-Bonding Corrections for Semiempirical QM Methods and Force Fields. J. Chem. Theroy Comput. 2010, 6, 38083816,  DOI: 10.1021/ct100408b
    29. 29
      Řezáč, J.; Hobza, P. Advanced Corrections of Hydrogen Bonding and Dispersion for Semiempirical Quantum Mechanical Methods. J. Chem. Theory Comput. 2012, 8, 141151,  DOI: 10.1021/ct200751e
    30. 30
      Stewart, J. J. P. Optimization of Parameters for Semiempirical Methods VI: More Modifications to the NDDO Approximations and Re-Optimization of Parameters. J. Mol. Model. 2013, 19, 132,  DOI: 10.1007/s00894-012-1667-x
    31. 31
      Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé-Hoover Chains: The Canonical Ensemble via Continuous Dynamics. J. Chem. Phys. 1992, 97, 26352643,  DOI: 10.1063/1.463940
    32. 32
      Flyvbjerg, H.; Petersen, H. G. Error Estimates on Averages of Correlated Data. J. Chem. Phys. 1989, 91, 461466,  DOI: 10.1063/1.457480