Hydrogen Bonding Environment of the N3–H Group of Flavin

May 22, 2017 - The light oxygen voltage (LOV) domain is a flavin-binding blue-light receptor domain, originally found in a plant photoreceptor phototr...
2 downloads 9 Views 953KB Size
Article

Hydrogen Bonding Environment of the N3–H Group of FMN in the LOV domains of Phototropins Tatsuya Iwata, Dai Nozaki, Atsushi Yamamoto, Takayuki Koyama, Yasuzo Nishina, Kiyoshi Shiga, Satoru Tokutomi, Masashi Unno, and Hideki Kandori Biochemistry, Just Accepted Manuscript • Publication Date (Web): 22 May 2017 Downloaded from http://pubs.acs.org on May 27, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Biochemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Hydrogen Bonding Environment of the N3–H Group of FMN in the LOV Domains of Phototropins Tatsuya Iwata,†,# Dai Nozaki,† Atsushi Yamamoto,† Takayuki Koyama,† Yasuzo Nishina,‡ Kiyoshi Shiga,§ Satoru Tokutomi,|| Masashi Unno,⊥ and Hideki Kandori*,† †

Department of Life Science and Applied Chemistry, Nagoya Institute of Technology, Showaku, Nagoya 466-8555, Japan ‡

Department of Molecular Physiology, Graduate School of Medical Sciences, Kumamoto University, Honjo, Kumamoto 860-8556, Japan

§

Department of Physiology, School of Health Sciences, Kumamoto University, Kuhonji, Kumamoto 862-0976, Japan

||

Department of Biological Science, Graduate School of Science, Osaka Prefecture University, Sakai, Osaka 599-8531, Japan ⊥

Department of Chemistry and Applied Chemistry, Graduate School of Science and Engineering, Saga University, Saga 840-8502, Japan

ACS Paragon Plus Environment

1

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 37

Abstract The light-oxygen-voltage (LOV) domain is a flavin-binding blue-light receptor domain, originally found in a plant photoreceptor phototropin (phot). Recently, LOV domains have been used in optogenetics as the photosensory domain of fusion proteins. Therefore, it is important to understand how LOV domains exhibit light-induced structural changes for the kinase domain regulation, which enables the design of LOV-containing optogenetics tools with higher photoactivation efficiency. In this study, the hydrogen bonding environment of the N3–H group of FMN of LOV2 domain from Adiantum neochrome (neo) 1 were investigated by lowtemperature Fourier transform infrared spectroscopy. Using specifically 15N-labeled FMN, [1,315

N2]FMN, the N3–H stretch was identified at 2831 cm–1 for the unphotolyzed state at 150 K,

indicating that the N3–H group forms a fairly strong hydrogen bond. The N3–H stretch showed temperature-dependence, with a shift to lower frequencies at ≤200 K and to higher frequencies at ≥250 K from the unphotolyzed to the intermediate states. Similar trends were observed in the LOV2 domains from Arabidopsis phot1 and phot2. By contrast, the N3–H stretch of Q1029L mutant of neo1-LOV2 and neo1-LOV1 were not temperature dependent in the intermediate state. These results seemed correlated with our previous finding that the LOV2 domains show the structural changes in the β-sheet region and/or the adjacent Jα helix of LOV2 domain but that such structural changes do not take place in Q1029L mutant or neo1-LOV1 domain. The environment around the N3–H group was also investigated.

ACS Paragon Plus Environment

2

Page 3 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Introduction Flavins are utilized as chromophores in some types of photoreceptors in organisms.1 The lightoxygen-voltage (LOV) domain is a flavin-binding photoreceptive domain, which was originally found in the photoreceptor domains of phototropin (phot),2, 3 a plant blue-light receptor. The LOV domain is named after the homology of the primary2 and tertiary4 structures to bacterial light-sensor photoactive yellow protein,5 oxygen sensor FixL,6 and the N-terminal domain of a voltage-gated potassium channel, HERG,7 the protein fold of which belongs to the Per-Arnt-Sim (PAS)L superfamily.8

Phot is a photoreceptor in plants where it is related to phototropic

response,9 relocation of chloroplasts10,

11

and stomata opening,12 and other functions. Phot

comprises about 1000 amino acid residues and two flavin mononucleotide (FMN) molecules, where the two LOV domains and FMN are at the N-terminal side and the Ser/Thr kinase domain is at the C-terminal side2. In addition to phots, LOV domains are found in other blue-light photoreceptors such as FKF113 and aureochrome.14 The LOV domains located at the N-terminal side regulate functional domains at the C-terminus, such as in phots, but vice versa in aureochromes. The two LOV domains have different functions although they have similar amino acid sequences.

The kinase activity is primarily regulated by the photoreaction of the LOV2

domain.15, 16 The LOV1 domain is referred to as the dimerization17, 18 and light attenuation site.16 Therefore, differences in the structure and structural changes in both LOV domains are of interest. The X-ray crystal structures of the LOV1 and LOV2 domains are quite similar, and there is no explanation for why the two LOV domains have different roles based on structural features.4, 19-21 The approximate positions of the LOV2 and kinase domains were determined by

ACS Paragon Plus Environment

3

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

X-ray small angle scattering22,

23

Page 4 of 37

in order to estimate how the LOV domain(s) regulates the

activity of the kinase domain. In 2003, an NMR study demonstrated the existence of an α-helix at the C-terminus of LOV2 and the α-helix regulates kinase activity.24 Structural changes in the α-helix were also studied using other techniques, including transient grating25-28 and Fourier transform infrared (FTIR) spectroscopy.29, 30 Our previous FTIR study showed that the structural changes in the Jα helix (or the corresponding region at the C-terminus of the LOV2 domain) differ among the LOV2 domains in Arabidopsis phot1 and phot2, and Adiantum neochrome 1 (neo1).30 Previous results suggest that the structural changes in Jα are specific to phot1 and phot2. The structural changes may differ between LOV1 and LOV2, and among the LOV2 domains of phot1, phot2, and neo1, but the photochemical reaction is generally the same in the LOV domains, where an adduct forms between FMN and a nearby cysteine.31-34 FMN absorbs light and after a singlet excited state of FMN, intersystem crossing occurs over a nanosecond timescale.35 Next, a shorter wavelength product appears with a time constant of the microsecond timescale,33, 36 which is called S390; this is the adduct formed between FMN and a cysteine. A recent study reported the existence of a pathway from the singlet excited state to the adduct form in the LOV2 domain of Chlamydomonas phot.37 S390 reverts to the unphotolyzed state in the dark within seconds to hours, where the time varies among LOV domains.31,

38, 39

Various

models have been proposed for the reactive cysteine, but previous FTIR studies indicate that the cysteine is protonated in both the unphotolyzed39-42 and the triplet excited43 states of FMN. Dynamic protein motion probably plays an important role in adduct formation.44 Recently, LOV domains have been utilized as photosensory domains in fusion proteins for use in optogenetics or light manipulation.45-47 Optogenetics is a technique for manipulating neurons

ACS Paragon Plus Environment

4

Page 5 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

or other cells in organisms by light illumination, where light-gated ion channels (channelrhodopsin) or other light-reactive enzymes are expressed. Therefore, it is important to determine how LOV domains exhibit light-induced structural changes in order to understand the appropriate regulation of these functional domains in optogenetics tools. The light-dependent regulation of a LOV domain as a functional domain in a fusion protein was demonstrated previously based on the α-helical movement of Jα helix in the LOV2 domain.47 However, the different structural changes of LOV2 domains suggest that diverse regulation systems may be employed in various types of enzymes.

Thus, some guiding principles are required for

optogenetics tools when designing fusion proteins with LOV domains and catalytic domains with various structures. In this study, in order to understand how the photochemical reaction of the FMN chromophore induces structural changes in LOV domains, we investigated the hydrogen bonding environments of FMN by FTIR spectroscopy. FMN forms hydrogen bonds with apoprotein via its C2=O, C4=O, and N3–H groups. We focused on the hydrogen bonds of N3–H stretch because we previously determined the hydrogen bonding environments of two C=O stretches in the FMN chromophore of Adiantum neo1-LOV2.48

We identified multiple temperature-dependent

structures for the S390 intermediate using low temperature FTIR spectroscopy in various LOV domains, including mutants.39 15

49-52

Thus, we prepared neo1-LOV2 reconstituted with [1,3-

N2]FMN in order to identify N3–H stretch and to investigate how the N3–H group changes its

hydrogen bonding environment. N3–H stretch was assigned to 2831 (–)/2812 (+) cm–1 measured at 150 K. We also identified the temperature dependence of the structural changes in N3–H stretch and the incapacity for the exchange of hydrogen for deuteron under D2O conditions. Similar changes were observed in Arabidopsis phot1-LOV2 and phot2-LOV2 despite the

ACS Paragon Plus Environment

5

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 37

presence of Jα helix. However, these changes in the hydrogen bonding strength were not observed in a Q1029L mutant, dehydrated neo1-LOV2, and neo1-LOV1, which did not exhibit progressive structural changes.39, 50 51 The hydrogen bonding environment of the N3–H group was strongly correlated with conformational changes in the peptide backbone.

Materials and Methods Preparation of Various LOV2 Domains.

Adiantum neo1-LOV2 (Pro905–Pro1087) and its

Q1029L mutant, and neo1-LOV1 (Gly660-Val805) were expressed as a fusion protein with calmodulin-binding peptide at the N-terminal side in Escherichia coli strain BL21(DE3) and purified.48-50

Arabidopsis phot1-LOV2-core (Lys462–Arg586), phot1-LOV2-Jα (Lys462–

Asp617), phot1-LOV2-Jα phot2-LOV2-core (Gln376–Gln500), phot2-LOV2-Jα (Gln376– Asp531), Adiantum neo1-LOV2-core (Arg916–Pro1045), and neo1-LOV2-Jα (Arg916– Asp1076) were expressed as a fusion protein with glutathione S-transferase (GST) at the Nterminal side in BL21(DE3), and the GST tag was removed by digestion with thrombin.29, 30 Specific site-directed 13C- and

15

N-labeled FMNs, i.e., [1,3-15N2]FMN, [4,10a-13C2]FMN, and

[2-13C]FMN, were synthesized as described previously.53 Reconstitution of FMN and apoprotein was performed as described previously.48 FTIR Measurements. FTIR measurements were obtained as described previously.48, 49 Infrared spectra of the hydrated films were measured using an FTS-7000 (DIGILAB) spectrophotometer. Hydrated films were illuminated with light at >400 nm using a combination of a halogentungsten lamp (1 kW) and a long-pass filter (L42, Asahi Techno Glass). The accumulated intermediate was only one kind, S390 intermediate, measured by UV-vis spectroscopy.40,

49

Spectral intensity was normalized by use of vibrational bands at 1500–1200 cm–1.54

ACS Paragon Plus Environment

6

Page 7 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Density Functional Theory (DFT) Calculations. The frequencies of N3–H stretch in 7,8dimethyl-10-glycerylisoalloxazine and its C4a–thiol adduct form with and without hydrogen bonding partners were determined from previous DFT calculations.54

The harmonic and

anharmonic vibrational frequencies were calculated using the Gaussian09 program.55

Results 15

N Isotope Effect of FMN on Vibrations at 1800–950 cm−1. In order to identify the vibration

signals where the N3 atom of FMN was involved, we prepared reconstituted neo1-LO2 with [1,3-15N2]FMN. Figure 1 shows the difference FTIR spectra for neo1-LOV2 reconstituted with unlabeled FMN (solid line) and [1,3-15N2]FMN (dotted line) measured at 150 (a) and 295 K (b) in the 1800–950 cm−1 region. At both temperatures, the identical bands exhibited downshifts of 4–10 cm−1, i.e., negative bands at 1507, 1407, 1272, and 1250 cm−1, and positive bands at 1541, 1521, 1379, 1258, and 1197 cm−1 at 150 K. Previously, we attributed the bands at 1551 (−)/1541 (+) and 1521 (+)/1507 (−) cm−1 to C10a=N1 stretching modes.48 However, the negative band at 1551 cm–1 did not exhibit isotope shift. According to the DFT calculations for 7,8-dimethyl-10-glycerylisoalloxazine,54 C10a=N1 stretch in the unphotolyzed and intermediate states were assigned to 1568/1512 and 1542 cm−1, respectively, and the band at 1551 cm−1 was assigned to C=C stretching involving ring I and C8a (see the inset in Figure 1 for details of the atom and ring numbering).

ACS Paragon Plus Environment

7

Biochemistry

1541

Difference Absorbance (0.01 abs./div.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 37

1521

1507

1379

1258

1197

(a) 150 K

1407 1272

1250

1551 1538 1519

1380 1195

1261

1504

1405

(b) 295 K

1272 1249

1553

1800

1600

1400

1200

1000

-1

Wavenumber (cm ) Figure 1. (a) Difference FTIR spectra of unlabeled neo1-LOV2 (solid line) and uniformly

15

N-labeled

(15N-FMN/15N-apoprotein) neo1-LOV2 (dotted line) in the 1800–950 cm−1 region measured at 150 K. (b) Difference FTIR spectra of neo1-LOV2 reconstituted with unlabeled FMN (solid line) and 15N-labeled (dotted line) apoprotein in the 1800–950 cm−1 region at 150 K. The inset shows the positions of the labeled nitrogen atoms in the isoalloxazine ring.

ACS Paragon Plus Environment

8

Page 9 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

The negative bands at 1407, 1272, and 1250 cm−1 exhibited isotope shifts of 4–7 cm−1. The lower frequencies suggest that these vibrations contained the single band characteristics of FMN, such as C2–N1, C2–N3, and C4–N3 stretching, and N3–H bending modes. The positive bands at 1379, 1258, and 1197 cm−1 exhibited isotope shifts of 5–9 cm−1. These bands are mainly attributable to ring stretching vibrations due to displacement of the N1 and N3 atoms. The bands at 1379 (+), 1272 (−), 1258 (+), and 1250 (−) exhibited spectral downshift for [4,10a-13C2]FMN and [2-13C]FMN,48 so they may be attributable to C4–N3, C2–N1, and C2–N3 stretches. According to DFT calculations by Kikuchi et al., these bands originate from C–N stretches in rings II and III of the isoalloxazine ring.54

The N3–H bending modes were calculated at

1443/1436 and 1445/1441 cm−1 in the presence of a hydrogen bonding acceptor for the unphotolyzed and S390 intermediate states, respectively. It is likely that the small frequency shift of the N3–H bending mode between the unphotolyzed and intermediate states prevented us from detecting the FTIR bands of N3–H bending in the difference FTIR spectra.

Identification of the N3–H Stretching Vibrational Mode. The higher frequency region was investigated to identify N3–H stretch in the FTIR spectra.

N3–H stretch of riboflavin

tetraacetate appeared at 3380 and 3200 cm−1 in chloroform solution, which are the hydrogen bonding-free and hydrogen-bonded forms, respectively.56 However, we found N3–H stretch in a much lower frequency region than 3200 cm−1. Figure 2 shows the difference FTIR spectra in the 3000–2740 cm−1 region where C–H stretches mainly appear. The peaks are as follows: 2970 (+), 2964 (–), 2949 (+), 2936 (−), 2929 (+), 2905 (+), 2878 (−), 2873 (+), 2864 (+), 2857 (−), 2831 (−), and 2812 (+) cm−1 (dotted lines in Figure 2a–e). It should be noted that both spectra agreed very well with each other for the unlabeled FMN-reconstituted and unreconstituted (native

ACS Paragon Plus Environment

9

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 37

formation of holoprotein in bacteria) LOV2 domains in this frequency region (Figure S1). In this region, the band at 2831 (−)/ 2811 (+) cm−1 exhibited downshift to 2816 (−)/2799 (+) cm−1 for [1,3-15N2]FMN-reconstituted noe1-LOV2 (Figure 2a). Bands appeared at 2816 (−)/2799 (+) cm−1 and at 2831 (−)/2811 (+) cm−1 for the uniformly 15N-labeled neo1-LOV2 (Figure 2b) and the reconstituted neo1-LOV2 of

14

N-FMN/15N-apoprotein (Figure 2c), respectively.

These

results indicate that the bands originate from the N3–H stretch of FMN. We could not exclude the possibility that a C–H group was influenced by 15N-labeling because C–H stretching modes appear in this region. Figure 2d and e show the difference spectra for 13Clabeled neo1-LOV2 and 12C-FMN/13C-apoprotein, respectively. The negative band at 2831 cm−1 exhibited downshift of 2 cm−1 for the

13

C-labeled neo1-LOV2 (Figure 2d), while a band

appeared at 2831 cm−1 for the 12C-FMN/13C-apoprotein (Figure 2e). The positive band at 2811 cm−1 did not exhibit downshift. The downshift of that of

15

13

C-labeled neo1-LOV2 was different from

N-labeled neo1-LOV2, so the bands at 2831 (−)/2811 (+) cm−1 originated from N–H

stretch rather than C–H stretch. The downshift of the N3–H stretching vibrational bands by 13C-labeling was affected indirectly by the

13

C-labeling of FMN.

In fact, the band downshifted by 2–3 cm−1 for neo1-LOV2

reconstituted with [4,10a-13C2]FMN and [2-13C2]FMN (Figure 2f and g, Table 1). These results show that 13C-labeling of the atoms neighboring FMN influenced N3–H stretch. The other bands in the 2970–2857 cm−1 region were downshifted by 4–10 cm−1 for the uniformly

13

C-labeled

neo1-LOV2, thereby indicating that these bands originated from C–H stretching modes (Figure 2d). These bands also downshifted for the

12

C-FMN/13C-apoprotein (Figure 2e), so they were

due to apoprotein and the C–H groups from the amino acid side chains.

ACS Paragon Plus Environment

10

Page 11 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Figure 2. Difference FTIR spectra for unlabeled (dotted lines in a, c, e–g), native (broken lines in b, and d), and labeled (solid lines) neo1-LOV2 in the 3000–2740 cm−1 region measured at 150 K. (a) [1,315

13

N2]FMN reconstituted, (b) uniformly

C-labeled, (e)

12

15

N-labeled, (c)

14

N-FMN/15N-apoprotein, (d) uniformly

C-FMN/13C-apoprotein, (f) [4,10a-13C2]FMN reconstituted, and (g) [2-

13

C]FMN reconstituted neo1-LOV2 are shown.

ACS Paragon Plus Environment

11

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 37

Table 1. Observed and calculated N3–H stretch in the unphotolyzed state.

Experimental

a, b

DFT calculationc, d model 1

model 2

2831 (–15, –3, –2) (150 K)

harmonic approximation

3473 (–8, 0, 0)

3114 (–8, 0, 0)

2821 (–18, –5, –7) (295 K)

anharmonic calculation

3447

2972 (–6, –2, 3)

model 3

model 4f

intermediate state

Experimental

a, b

DFT calculationc, d

3183 (–2, 0, 0) harmonic approximation

3477 (–8, 0, 0)

2812 (–13, –2, n.d.e)

3138 (–6, 0, 0)

2839 (–7, –5, –5)

3081 (–9, 1, –5) anharmonic calculation

3352 3035 (–8, –5, –7)

a

The numbers in parentheses are the isotope shifts for [1,3-15N2]FMN, [4,10a-13C2]FMN, and

[2-13C]FMN, respectively. bThe top and bottom frequencies are those measured at 150 and 295 K, respectively. cThe numbers in parentheses are the isotope shifts for [1,3-15N2]-na, [4,10a13

C2]-na, and [2-13C]-na, respectively. dThe model structures for the DFT calculations were

obtained from Kikuchi et al.54

DFT Calculation of N3–H Stretch in the Absence and Presence of Hydrogen Bonding Partners. DFT calculations were performed to investigate the effects of hydrogen bonding around the N3– H group of FMN. According to previous DFT calculations for N3–H stretch in 7,8-dimethyl-10glycerylisoalloxazine, N–H stretch was identified at 3473 and 3115 cm−1 in the absence and presence of a hydrogen bonding partner, respectively, i.e., the C=O group of Asn998 (models 1 and 2 given by Kikuchi et al.,54 Table 1). This suggests the existence of a strong hydrogen bond

ACS Paragon Plus Environment

12

Page 13 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

at the N3–H group due to the presence of hydrogen bonding acceptor(s) because the N–H and O– H stretch in the lower frequency region indicates the formation of hydrogen bonds. In addition, it has been reported that large anharmonic effects appear in X–H (X = O, N, C) stretches,57 thereby indicating that the observed lower frequency of N3–H stretch may have originated from anharmonic effects rather than hydrogen bonding effects. Therefore, a DFT calculation that considered the anharmonicity was performed (Table 1). The anharmonic frequency of N3–H stretch for 7,8-dimethyl-10-glycerylisoalloxazine was calculated at 3447 and 2972 cm−1 in the absence and presence of hydrogen bonding acceptors, respectively. A lower frequency shift was also shown by the harmonic calculation. The differences between the harmonic and anharmonic frequencies were 26 and 143 cm−1 in the absence and presence of hydrogen bonding acceptors, respectively. The frequency at 2972 cm−1 is much closer to the observed value (2831 cm−1). This suggests that the lower frequency shift of N3–H stretch also originated from the increased anharmonicity as well as from hydrogen bonding formation. In the adduct form of the model compound (model 3 given by Kikuchi et al.54), N3−H stretch appeared at 3477 and 3352 cm−1 without the hydrogen bonding partners based on calculations of the harmonic approximation and anharmonic contribution, respectively. In the adduct form, the anharmonic effect might not be negligible because the frequency difference at 125 cm−1 was larger than the frequency difference for the non-adduct form. With a hydrogen bonding donor, the N3–H group exhibited vibrational coupling with the N–H group from Asn980, thereby resulting in two frequencies of N3–H stretch, i.e., 3183 and 3139 cm−1, and 3081 and 3035 cm−1. Thus, the N3–H stretch appeared in a region at a frequency that was ~300 cm−1 lower in the presence of the hydrogen bonding acceptors.

ACS Paragon Plus Environment

13

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The isotope effect of N3–H stretch due to

Page 14 of 37

13

C labeling of nearby carbon atoms in FMN was

identified by the anharmonic calculation but not by the harmonic calculation (Table 1). The trend in the calculated frequency shift was similar to that measured, which indicates that the anharmonicity is not negligible.

H/D Exchange of the N3–H Group. Figure 3 shows the difference FTIR spectra for hydrated films prepared with H2O (solid lines) and D2O (dotted lines) in the 3000–2740 (a) and 1800–950 (b) cm−1 regions at 150 K. When the N–H group was deuterated, the band at 2831 cm−1 was expected to appear at ~2070 cm−1. However, the band at 2831 (−)/2812 (+) cm−1 remained in these conditions (Figure 3a) and no band was observed in the 2250–2000 cm–1 region (see Figure 2 in Nozaki et al.58). Thus, the N3–H group was not deuterated by the hydration of D2O in neo1LOV2. In our experimental conditions, H/D exchange was moderate so the unchanged X–H groups formed strong hydrogen bonds in proteins. As shown above, it is reasonable to assume that N3–H formed a very strong hydrogen bond so H/D exchange did not occur. In the 1800–950 cm−1 region, an H/D exchange effect was observed in the 1660–1620 and 1300–1250 cm−1 regions and the 1084 (−) cm−1 band. The N3–H group was not deuterated, so this effect originated from the deuteration of the exchangeable X–H groups in the peptide backbone (N–H groups) and of the side chains (N–H and O–H groups). The decreases in the intensities of the bands at 1660 (+), 1645 (−), and 1627 (+) cm−1, and the appearance of the band at 1639 (+) cm−1 are attributable to the amide-I vibrational mode, which was affected by the N–D groups of the peptide backbone. The spectral differences at 1300–1250 cm−1 are attributable to the amide-III vibrational mode.

ACS Paragon Plus Environment

14

Page 15 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Figure 3. Difference FTIR spectra for the unlabeled neo1-LOV2 hydrated with H2O (solid lines) and D2O (dotted lines) in the 3000–2740 cm−1 region (a) and 1800–950 cm−1 region (b) measured at 150 K.

Swartz et al. showed that there was an H/D effect according to the difference FTIR spectra obtained for oat phot1-LOV2 in the 1300–1200 cm−1 region.59 They suggested that the N3–H group is accessible to bulk protons via protein residues or structural water molecules based on their experimental pH-dependent titration of fluorescence.33 In the case of neo-LOV2, the shifts were not due to deuteration of the N3–H group, although structural waters were exchanged for D2O.58 It is unlikely that H/D exchange can occur readily for the N3–H group in LOV domains. Ataka et al. suggested the possibility of asymmetric P–O stretches in the phosphate group of

ACS Paragon Plus Environment

15

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 37

FMN based on their study of Chlamydomonas phot-LOV1.41 The origin of the negative band at 1084 cm–1 is unclear because it disappeared or shifted elsewhere.

Temperature Dependence of the Vibrations in the 3000–2740 cm−1 Region. Difference FTIR spectra were measured at 200, 250, and 295 K in addition to 150 K (Figure 4), where different structural changes in the peptide backbone were observed despite the same chromophore.48, 49 As shown in Figure 4a (reproduced from Figure 2a), at 150 K, the band for N3–H stretch appeared at 2831 (−)/2812 (+) cm−1. A similar signal was also observed at 2828 (−)/2805 (+) cm−1 at 200 K (Figure 4b). N3–H stretch of the S390 intermediate state appeared at lower frequencies compared with that of the unphotolyzed state. Moreover, N3–H stretch shifted to higher frequencies after the formation of S390 at 250 and 295 K, i.e., to 2843 (+)/2824 (−) and 2839 (+)/2821 (–) cm−1, respectively (Figure 4c and d). The hydrogen bonds of the N3–H group in the S390 state were weaker at ≥250 K than at ≤200 K. The structural change in FMN (FMN-cysteinyl adduct) did not occur between 200 and 250 K for the S390 intermediate, so the change in the N3–H group was caused by apoprotein rather than by FMN itself. The hydrogen bonding environment of the N3–H group for the S390 state was temperature dependent. The C=O group of Asn998 (for Adiantum neo1-LOV2) is a partner in the hydrogen bond of the N3–H group, so the environment around Asn998 must be changed. Asn998 comprises a β-strand and the transition temperature of the structural changes in the N3– H group corresponded to that at which the structural changes in the β-sheet region appeared. Therefore, the hydrogen bonding strength of the N3–H group reflected the structural change in Asn998.

ACS Paragon Plus Environment

16

Page 17 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

A temperature-dependent effect was also observed in the 3000–2850 cm–1 region, which originated from apoprotein (see Figure 2d and e). The spectrum obtained at 200 K was similar to that at 150 K, but the spectra were different at 250 and 295 K. The bands at 2979 (+), 2958 (–), 2949 (+), 2920 (–) 2876 (+), and 2860 (–) cm−1 at 295 K were also assigned to the C–H stretching vibrations of apoprotein based on the uniformly reconstituted neo1-LOV2 of

12

C-FMN/13C-apoprotein.

13

C-labeled neo1-LOV2 and

We determined the temperature

dependence of the amide-I region, which exhibited progressive structural changes in the peptide backbone and the C–H stretching region also reflected the differences in the structural changes with temperature.

ACS Paragon Plus Environment

17

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 37

Figure 4. Difference FTIR spectra for neo1-LOV2 reconstituted with unlabeled (solid lines) and [1,3-15N2] (dotted lines) FMN in the 3000–2740 cm–1 region measured at 150 (a), 200 (b), 250 (c), and 295 K (d).

ACS Paragon Plus Environment

18

Page 19 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Correlation between the Hydrogen Bond of N3–H and Structural Changes in the β-sheet in the S390 Intermediate. Previously, we reported that the unhydrated wild type (WT) neo1-LOV2 and Q1029L mutant did not exhibit structural changes in the β-sheet region after the formation of S390 at 295 K, whose structural changes were similar as those of WT neol-LOV2 at lower temperatures.50, 51 Thus, we expected that N3–H stretch would appear at a lower frequency in S390 for these samples as is the case for the measurement of WT at ≤200 K (see Figure 4a and b). Figure 5 shows the difference FTIR spectra for the unhydrated WT (a) and hydrated Q1029L mutant (b) reconstituted with unlabeled (solid lines) and [1,3-15N2]FMN (dotted lines). In this mutant, at 295 K, N3–H stretch appeared in the 2824(–)/2795 (+) and 2834(–)/2814 (+) cm−1 regions for the unhydrated WT and hydrated Q1029L, respectively. The N3–H stretching mode appeared at a lower frequency in S390, which was similar to that determined for the hydrated WT measured at 150 and 200 K (see Figures 2 and 3). These results indicate that there was a correlation between the upshift of N3–H stretch and structural changes in the β-sheet region. Similar trends were observed in the FTIR spectra of the LOV2 domains with other amino acid lengths in Adiantum neo1-LOV2, and in Arabidopsis phot1 and phot2 (Figure S2a–c). Reconstitution of labeled FMN was not conducted, but the bands around 2800 cm–1 were temperature dependent between 150 and 295 K, where the former appeared at lower frequencies and the latter at higher frequencies. By contrast, the N3–H stretch in Adiantum neo1-LOV1 was not temperature dependent (Figure S2d). The hydrogen bonding environment of the N3–H group was very strong among the LOV domains, and the hydrogen bonding strength and structural changes in apoprotein were strongly correlated with each other.

ACS Paragon Plus Environment

19

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 37

Figure 5. Difference FTIR spectra for neo1-LOV2 reconstituted with unlabeled (solid lines) and [1,3-15N2] (dotted lines) FMN in the 1800–950 cm−1 region at 150 K (a) and 295 K (b).

Discussion Hydrogen Bonding Environment of the N3–H Group of FMN in Neo1-LOV2. In this study, we assigned the FTIR band for N3–H stretch in FMN and the structural changes after adduct formation in neo1-LOV2. The N3–H stretching band appeared at 2831 and 2821 cm−1 in the unphotolyzed state at 150 and 295 K, respectively. The frequency at which the band appeared

ACS Paragon Plus Environment

20

Page 21 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

was quite low for N3–H stretch in flavin compared with that in riboflavin tetraacetate in chloroform at 3380 cm–1.56 The low frequency shift of N3–H stretch was explained by the presence of the hydrogen bonding partner according to the DFT calculations for 7,8-dimethyl10-glycerylisoalloxazine. In fact, the H/D exchange of the N3–H group did not occur under mild deuteration conditions. Previous studies have reported that X–H (X = N, O) stretches appeared at lower frequencies (>3000 cm−1) in some photoreceptive proteins. Thus, the N–H stretch in the retinal protonated Schiff base of bacteriorhodopsin appeared at ~2800 cm−1 and ~3350 cm−1 in the ground state and K intermediate, respectively.60 The positively charged N–H group of the Schiff base interacted with the negatively charged counterion Asp85, so it was expected that the N–H group would form a strong hydrogen bond. The N–H stretch of histidine residue was also observed in the 2800–2600 cm−1 region (as a Fermi resonance signal) for photosystem II61,62 and sensory rhodopsin I.63 In the case of photosystem II, strongly hydrogen bonded histidines were observed sandwiched with Fe2+ and negatively charged quinone,61 or they coordinated with manganese ion in the oxygen evolving complex.62 In the case of sensory rhodopsin I, it is not known whether the positively charged histidine residue had a strong hydrogen bond via its hydrogen bonding partner.63 In both cases, charged species were involved in the strong hydrogen bonds. In BLUF domains, the O–H stretch due to a tyrosine residue appeared at ~2800 cm−1 in the red-shifted intermediate state, where there was quite strong hydrogen bonding, probably with a nearby glutamine residue.64 However, in this case, no charges were apparent in the specific hydrogen bonding environment. At present, it is not clear whether strong hydrogen bonds are formed with the N3–H group in the LOV domain, but this study of BLUF domains might suggest that they are.

ACS Paragon Plus Environment

21

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 37

The downshift of the N3–H group due to hydrogen bonding formation was estimated as ~500 cm−1. Dürr et al. reported that riboflavin can bind to oat phot1-LOV2 apoprotein and that the R40D mutant can bind riboflavin.65 Riboflavin does not have a phosphate group and Arg40 is an ionic bond partner of the phosphate group in FMN. Thus, the ionic bond between FMN and apoprotein is not necessarily essential. Our findings suggest that a specific hydrogen bond interaction between the N3–H group of flavin and the C=O group of Asn998 has important roles in the binding of flavin to apoprotein.

Structural Changes in the N3–H Group in the S390 State. The hydrogen bond of the N3–H group was strengthened at ≤200 K but weakened at ≥250 K in the S390 state (Figure 3). Because the accumulated intermediate is only S390, the formation of FMN-thiol adduct, at any temperature measured, this change can be explained not by the structural change of FMN itself but by the interaction between a hydrogen bond partner, which is the C=O group of Asn998.4 This indicates that the structural change in the N3–H group is an indicator of the structural change in Asn998. The simplest interpretation is that the distance between the N3–H group of FMN and the C=O group of Asn998 was closer at ≤200 K but more distant at ≥250 K. Figure 6 shows a schematic model of the hydrogen bonding environment between FMN and apoprotein, which includes our previously reported results for the C2=O and C4=O groups.48 N3–H forms a strong hydrogen bond in the unphotolyzed state (left panel of Figure 6), but the bond becomes much stronger in the S390 intermediate (middle panel of Figure 6). However, the hydrogen bonds weaken for the C2=O and C4=O groups. The crystal structures of the unphotolyzed and S390 states demonstrate that the distance between the oxygen atom of C2=O and the nitrogen atom of the Asn998 side chain increases, whereas the distance between the N3 nitrogen and the

ACS Paragon Plus Environment

22

Page 23 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

oxygen atom of the Asn998 side chain decreases.4, 34 In the case of neo1-LOV2, we proposed two states for S390, i.e., S390I and S390II, where S390I exhibits structural change in the turn and α-helix structures, and S390II exhibits structural changes in the β-sheet.66 The crystal structure of S390 probably represents S390I. The resolutions of the crystal structures are 2.7 and 2.6 Å in the unphotolyzed and S390 states, respectively, so the difference of 0.1 Å may not be sufficiently reliable. Our results support the observations based on the crystal structures. In the S390II state, the hydrogen bond of the N3–H group weakens.

The simplest

interpretation of this change is that the distance between Asn998 and FMN increases, and Asn998 probably moves far from FMN. Considering that Asn998 is located in the β-strand and that the structural change occurs in the β-sheet region, then Asn998 moves far from FMN. This strongly suggests that the β-sheet region responsible for the structural change is that containing Asn998. Previously, we showed the importance of Gln1029 for structural changes in the βsheet.50 Gln1029 switches the hydrogen bonding partner after adduct formation, which begins with structural changes in the β-sheet. The structural change of Gln1029, which is located on the βE strand, is then transferred to Asn998 via the βD strand. Phenomena similar to N3–H stretch in FMN have been observed in other LOV2 domains (Figure S2). It is curious that the apparent structural changes in the β-sheet were quite small for the LOV2-core domains in Arabidopsis phot1 compared with the LOV2-core domain in Adiantum neo1. However, the hydrogen bonding environment of the N3–H group is temperature dependent, thereby implying that structural changes occur in the β-sheet or β-strand where the asparagine (Asn544 for Arabidopsis phto1) is located. Thus, common mechanisms may explain the structural changes in LOV2 domains. All LOV2 domains exhibit temperature dependence in terms of the hydrogen bonding strength of the N3–H

ACS Paragon Plus Environment

23

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 37

group in the S390 state, which reflects the progressive structural changes in the β-sheet or βstrand where Asp544 is located. However, it is still unclear how the common structural changes in the N3–H group of FMN and the nearby asparagine induce different structural changes, i.e., in the Jα helix for Arabidopsis phto1 and in the β-sheet for neo1-LOV2, and thus the mechanisms responsible must be determined. Elucidating these mechanisms will help to understand the kinase domain activation mechanisms and guide the construction of fusion proteins that can be activated/deactivated by the photoreaction involving LOV domains.

Figure 6. Schematic showing the hydrogen bonding environment between FMN and apoprotein in neo1LOV2 for the unphotolyzed (PDB entry: 1G28) and S390 (PDB entry: 1JNU) states. The peptide backbone is represented as ribbons. FMN is shown as a stick drawing, while the side chains of Cys966, Asn998, Asn1008, and Gln1029 are shown as ball-and-stick drawings. Thick dotted lines represent hydrogen bonds and numbers show the O…N distances in angstroms. Thin dotted lines show the positions where the structural changes are expected to occur.

ACS Paragon Plus Environment

24

Page 25 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Supporting Information The following files are available free of charge. Difference FTIR spectra for native and uniformly 15N-labeled neo1-LOV2 in the 3000–2730 cm– 1

region (Figure S1, PDF). Difference FTIR spectra for Arabidopsis phot1-LOV2 and phot2-

LOV2, and Adiantum neo1-LOV2 and neo1-LOV1 in the 3000–2760 cm–1 region (Figure S2, PDF).

Corresponding Author * To whom correspondence should be addressed. Phone & FAX: 81-52-735-5207. E-mail: [email protected]

Present Addresses #

T. I.: Department of Pharmaceutical Science, Toho University, Funabashi, Chiba 274–8510,

Japan

Funding Sources This study was supported by grants from the Japanese Ministry of Education, Culture, Sports, Science, and Technology to H. K. (25104009, 15H02391), M. U. (26410017), and T. I. (16K07318).

Some of the computations were performed at the Research Center for

Computational Science, Okazaki, Japan.

ACS Paragon Plus Environment

25

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 37

Acknowledgements The authors would like to thank Enago (www.enago.jp) for the English language review.

ABBREVIATIONS LOV, light-oxygen-voltage; PAS, Per-Arnt-Sim; FMN, flavin mononucleotide; FTIR, Fourier transform infrared.

ACS Paragon Plus Environment

26

Page 27 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

REFERENCES [1] Losi, A., and Gartner, W. (2012) The evolution of flavin-binding photoreceptors: an ancient chromophore serving trendy blue-light sensors, Annu. Rev. Plant Biol. 63, 49–72. [2] Huala, E., Oeller, P. W., Liscum, E., Han, I. S., Larsen, E., and Briggs, W. R. (1997) Arabidopsis NPH1: a protein kinase with a putative redox-sensing domain, Science 278, 2120– 2123. [3] Christie, J. M., Salomon, M., Nozue, K., Wada, M., and Briggs, W. R. (1999) LOV (light, oxygen, or voltage) domains of the blue-light photoreceptor phototropin (nph1): binding sites for the chromophore flavin mononucleotide, Proc. Natl. Acad. Sci. U.S.A. 96, 8779–8783. [4] Crosson, S., and Moffat, K. (2001) Structure of a flavin-binding plant photoreceptor domain: insights into light-mediated signal transduction, Proc. Natl. Acad. Sci. U.S.A. 98, 2995– 3000. [5] Borgstahl, G. E., Williams, D. R., and Getzoff, E. D. (1995) 1.4 A structure of photoactive yellow protein, a cytosolic photoreceptor: unusual fold, active site, and chromophore, Biochemistry 34, 6278–6287. [6] Miyatake, H., Mukai, M., Park, S. Y., Adachi, S., Tamura, K., Nakamura, H., Nakamura, K., Tsuchiya, T., Iizuka, T., and Shiro, Y. (2000) Sensory mechanism of oxygen sensor FixL from Rhizobium meliloti: crystallographic, mutagenesis and resonance Raman spectroscopic studies, J. Mol. Biol. 301, 415–431.

ACS Paragon Plus Environment

27

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 37

[7] Morais Cabral, J. H., Lee, A., Cohen, S. L., Chait, B. T., Li, M., and Mackinnon, R. (1998) Crystal structure and functional analysis of the HERG potassium channel N terminus: a eukaryotic PAS domain, Cell 95, 649–655. [8] Takahashi, J. S. (1992) Circadian clock genes are ticking, Science 258, 238–240. [9] Liscum, E., and Briggs, W. R. (1995) Mutations in the NPH1 locus of Arabidopsis disrupt the perception of phototropic stimuli, Plant Cell 7, 473–485. [10] Kagawa, T., Sakai, T., Suetsugu, N., Oikawa, K., Ishiguro, S., Kato, T., Tabata, S., Okada, K., and Wada, M. (2001) Arabidopsis NPL1: a phototropin homolog controlling the chloroplast high-light avoidance response, Science 291, 2138–2141. [11] Sakai, T., Kagawa, T., Kasahara, M., Swartz, T. E., Christie, J. M., Briggs, W. R., Wada, M., and Okada, K. (2001) Arabidopsis nph1 and npl1: blue light receptors that mediate both phototropism and chloroplast relocation, Proc. Natl. Acad. Sci. U.S.A. 98, 6969–6974. [12] Kinoshita, T., Doi, M., Suetsugu, N., Kagawa, T., Wada, M., and Shimazaki, K. (2001) Phot1 and phot2 mediate blue light regulation of stomatal opening, Nature 414, 656–660. [13] Nelson, D. C., Lasswell, J., Rogg, L. E., Cohen, M. A., and Bartel, B. (2000) FKF1, a clock-controlled gene that regulates the transition to flowering in Arabidopsis, Cell 101, 331– 340. [14] Takahashi, F., Yamagata, D., Ishikawa, M., Fukamatsu, Y., Ogura, Y., Kasahara, M., Kiyosue, T., Kikuyama, M., Wada, M., and Kataoka, H. (2007) AUREOCHROME, a photoreceptor required for photomorphogenesis in stramenopiles, Proc. Natl. Acad. Sci. U.S.A. 104, 19625–19630.

ACS Paragon Plus Environment

28

Page 29 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

[15] Christie, J. M., Swartz, T. E., Bogomolni, R. A., and Briggs, W. R. (2002) Phototropin LOV domains exhibit distinct roles in regulating photoreceptor function, Plant J. 32, 205–219. [16] Matsuoka, D., and Tokutomi, S. (2005) Blue light-regulated molecular switch of Ser/Thr kinase in phototropin, Proc. Natl. Acad. Sci. U.S.A. 102, 13337–13342. [17] Salomon, M., Lempert, U., and Rudiger, W. (2004) Dimerization of the plant photoreceptor phototropin is probably mediated by the LOV1 domain, FEBS Lett. 572, 8–10. [18] Nakasako, M., Iwata, T., Matsuoka, D., and Tokutomi, S. (2004) Light-induced structural changes of LOV domain-containing polypeptides from Arabidopsis phototropin 1 and 2 studied by small-angle X-ray scattering, Biochemistry 43, 14881–14890. [19] Fedorov, R., Schlichting, I., Hartmann, E., Domratcheva, T., Fuhrmann, M., and Hegemann, P. (2003) Crystal structures and molecular mechanism of a light-induced signaling switch: The Phot-LOV1 domain from Chlamydomonas reinhardtii, Biophys J. 84, 2474–2482. [20] Halavaty, A. S., and Moffat, K. (2007) N- and C-terminal flanking regions modulate lightinduced signal transduction in the LOV2 domain of the blue light sensor phototropin 1 from Avena sativa, Biochemistry 46, 14001–14009. [21] Nakasako, M., Zikihara, K., Matsuoka, D., Katsura, H., and Tokutomi, S. (2008) Structural basis of the LOV1 dimerization of Arabidopsis phototropins 1 and 2, J. Mol. Biol. 381, 718–733. [22] Okajima, K., Aihara, Y., Takayama, Y., Nakajima, M., Kashojiya, S., Hikima, T., Oroguchi, T., Kobayashi, A., Sekiguchi, Y., Yamamoto, M., Suzuki, T., Nagatani, A., Nakasako, M., and Tokutomi, S. (2014) Light-induced conformational changes of LOV1 (light oxygen

ACS Paragon Plus Environment

29

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 37

voltage-sensing domain 1) and LOV2 relative to the kinase domain and regulation of kinase activity in Chlamydomonas phototropin, J. Biol. Chem. 289, 413–422. [23] Takayama, Y., Nakasako, M., Okajima, K., Iwata, A., Kashojiya, S., Matsui, Y., and Tokutomi, S. (2011) Light-induced movement of the LOV2 domain in an Asp720Asn mutant LOV2-kinase fragment of Arabidopsis phototropin 2, Biochemistry 50, 1174–183. [24] Harper, S. M., Neil, L. C., and Gardner, K. H. (2003) Structural basis of a phototropin light switch, Science 301, 1541–1544. [25] Eitoku, T., Nakasone, Y., Matsuoka, D., Tokutomi, S., and Terazima, M. (2005) Conformational dynamics of phototropin 2 LOV2 domain with the linker upon photoexcitation, J. Am. Chem. Soc. 127, 13238–13244. [26] Nakasone, Y., Eitoku, T., Matsuoka, D., Tokutomi, S., and Terazima, M. (2006) Kinetic measurement of transient dimerization and dissociation reactions of Arabidopsis phototropin 1 LOV2 domain, Biophys J. 91, 645–653. [27] Nakasone, Y., Eitoku, T., Matsuoka, D., Tokutomi, S., and Terazima, M. (2007) Dynamics of conformational changes of Arabidopsis phototropin 1 LOV2 with the linker domain, J. Mol. Biol. 367, 432–442. [28] Eitoku, T., Nakasone, Y., Zikihara, K., Matsuoka, D., Tokutomi, S., and Terazima, M. (2007) Photochemical intermediates of Arabidopsis phototropin 2 LOV domains associated with conformational changes, J. Mol. Biol. 371, 1290–1303.

ACS Paragon Plus Environment

30

Page 31 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

[29] Yamamoto, A., Iwata, T., Sato, Y., Matsuoka, D., Tokutomi, S., and Kandori, H. (2009) Light signal transduction pathway from flavin chromophore to the J alpha helix of Arabidopsis phototropin1, Biophys J. 96, 2771–2778. [30] Koyama, T., Iwata, T., Yamamoto, A., Sato, Y., Matsuoka, D., Tokutomi, S., and Kandori, H. (2009) Different role of the Jalpha helix in the light-induced activation of the LOV2 domains in various phototropins, Biochemistry 48, 7621–7628. [31] Salomon, M., Christie, J. M., Knieb, E., Lempert, U., and Briggs, W. R. (2000) Photochemical and mutational analysis of the FMN-binding domains of the plant blue light receptor, phototropin, Biochemistry 39, 9401–9410. [32] Salomon, M., Eisenreich, W., Durr, H., Schleicher, E., Knieb, E., Massey, V., Rudiger, W., Muller, F., Bacher, A., and Richter, G. (2001) An optomechanical transducer in the blue light receptor phototropin from Avena sativa, Proc. Natl. Acad. Sci. U.S.A. 98, 12357–12361. [33] Swartz, T. E., Corchnoy, S. B., Christie, J. M., Lewis, J. W., Szundi, I., Briggs, W. R., and Bogomolni, R. A. (2001) The photocycle of a flavin-binding domain of the blue light photoreceptor phototropin, J. Biol. Chem. 276, 36493–36500. [34] Crosson, S., and Moffat, K. (2002) Photoexcited structure of a plant photoreceptor domain reveals a light-driven molecular switch, Plant Cell 14, 1067–1075. [35] Kennis, J. T., Crosson, S., Gauden, M., van Stokkum, I. H., Moffat, K., and van Grondelle, R. (2003) Primary reactions of the LOV2 domain of phototropin, a plant blue-light photoreceptor, Biochemistry 42, 3385–3392.

ACS Paragon Plus Environment

31

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 37

[36] Kottke, T., Heberle, J., Hehn, D., Dick, B., and Hegemann, P. (2003) Phot-LOV1: photocycle of a blue-light receptor domain from the green alga Chlamydomonas reinhardtii, Biophys J. 84, 1192–1201. [37] Zhu, J., Mathes, T., Hontani, Y., Alexandre, M. T., Toh, K. C., Hegemann, P., and Kennis, J. T. (2016) Photoadduct Formation from the FMN Singlet Excited State in the LOV2 Domain of Chlamydomonas reinhardtii Phototropin, J. Phys. Chem. Lett. 7, 4380–4384. [38] Kasahara, M., Swartz, T. E., Olney, M. A., Onodera, A., Mochizuki, N., Fukuzawa, H., Asamizu, E., Tabata, S., Kanegae, H., Takano, M., Christie, J. M., Nagatani, A., and Briggs, W. R. (2002) Photochemical properties of the flavin mononucleotide-binding domains of the phototropins from Arabidopsis, rice, and Chlamydomonas reinhardtii, Plant Physiol. 129, 762– 773. [39] Iwata, T., Nozaki, D., Tokutomi, S., and Kandori, H. (2005) Comparative investigation of the LOV1 and LOV2 domains in Adiantum phytochrome3, Biochemistry 44, 7427–7434. [40] Iwata, T., Tokutomi, S., and Kandori, H. (2002) Photoreaction of the cysteine S-H group in the LOV2 domain of adiantum phytochrome3, J. Am. Chem. Soc. 124, 11840–11841. [41] Ataka, K., Hegemann, P., and Heberle, J. (2003) Vibrational spectroscopy of an algal Phot-LOV1 domain probes the molecular changes associated with blue-light reception, Biophys J. 84, 466–474. [42] Bednarz, T., Losi, A., Gartner, W., Hegemann, P., and Heberle, J. (2004) Functional variations among LOV domains as revealed by FT-IR difference spectroscopy, Photochem. Photobiol. Sci. 3, 575–579.

ACS Paragon Plus Environment

32

Page 33 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

[43] Sato, Y., Iwata, T., Tokutomi, S., and Kandori, H. (2005) Reactive cysteine is protonated in the triplet excited state of the LOV2 domain in Adiantum phytochrome3, J. Am. Chem. Soc. 127, 1088–1089. [44] Dozaki, D., Iwata, T., Tokutomi, S., and Kandori, H. (2005) Unique temperature dependence in the adduct formation between FMN and cysteine S-H group in the LOV2 domain of Adiantum phytochrome3, Chem. Phys. Lett. 410, 59–63. [45] Strickland, D., Moffat, K., and Sosnick, T. R. (2008) Light-activated DNA binding in a designed allosteric protein, Proc. Natl. Acad. Sci. U.S.A. 105, 10709–10714. [46] Lee, J., Natarajan, M., Nashine,V. C., Socolich, M., Vo, T., Russ, W. P., Benkovic, S. J., and Ranganathan, R. (2008) Surface sites for engineering allosteric control in proteins, Science 322, 438–442. [47] Wu, Y. I., Frey, D., Lungu, O. I., Jaehrig, A., Schlichting, I., Kuhlman, B., and Hahn, K. M. (2009) A genetically encoded photoactivatable Rac controls the motility of living cells, Nature 461, 104–108. [48] Iwata, T., Nozaki, D., Sato, Y., Sato, K., Nishina, Y., Shiga, K., Tokutomi, S., and Kandori, H. (2006) Identification of the C=O stretching vibrations of FMN and peptide backbone by

13

C-labeling of the LOV2 domain of Adiantum phytochrome3, Biochemistry 45, 15384–

15391. [49] Iwata, T., Nozaki, D., Tokutomi, S., Kagawa, T., Wada, M., and Kandori, H. (2003) Light-induced structural changes in the LOV2 domain of Adiantum phytochrome3 studied by low-temperature FTIR and UV-visible spectroscopy, Biochemistry 42, 8183–8191.

ACS Paragon Plus Environment

33

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 37

[50] Nozaki, D., Iwata, T., Ishikawa, T., Todo, T., Tokutomi, S., and Kandori, H. (2004) Role of Gln1029 in the photoactivation processes of the LOV2 domain in Adiantum phytochrome3, Biochemistry 43, 8373–8379. [51] Iwata, T., Yamamoto, A., Tokutomi, S., and Kandori, H. (2007) Hydration and temperature similarly affect light-induced protein structural changes in the chromophoric domain of phototropin, Biochemistry 46, 7016–7021. [52] Yamamoto, A., Iwata, T., Tokutomi, S., and Kandori, H. (2008) Role of phe1010 in lightinduced structural changes of the neol-LOV2 domain of Adiantum, Biochemistry 47, 922–928. [53] Miura, R., Nishina, Y., Ohta, M., Tojo, H., Shiga, K., Watari, H., Yamano, T., and Miyake, Y. (1983) Resonance Raman study on the flavin in the purple intermediates of D-amino acid oxidase, Biochem. Biophys. Res. Commun. 111, 588–594. [54] Kikuchi, S., Unno, M., Zikihara, K., Tokutomi, S., and Yamauchi, S. (2009) Vibrational assignment of the flavin-cysteinyl adduct in a signaling state of the LOV domain in FKF1, J. Phys. Chem. B 113, 2913–2921. [55] Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery, J. A. Jr, Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, N. J., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E.,

ACS Paragon Plus Environment

34

Page 35 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

Yazyev, O., Austin, A. J., Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P. Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, O., Foresman, J. B., Ortiz, J. V., Cioslowski, J., and Fox, D. J. Gaussian 09, Rev. D.01, Gaussian, Inc., Wallingford, CT, USA, 2009. [56] Kyogoku, Y., and Yu, B. S. (1968) The specific hydrogen bonding of riboflavin derivatives to an adenine compound, Bull. Chem. Soc. Japan 41, 1742. [57] Daněček, P., Kapitán, J., Baumruk, V., Bednárova, L., Kopecky, V., Jr., Bouř, P. (2007) Anharmonic effects in IR, Raman, and Raman optical activity spectra of alanine and proline zwitterions. J. Chem. Phys. 126, 224513. [58] Nozaki, D., Iwata, T., Tokutomi, S., and Kandori, H. (2005) Water structural changes in the activation process of them LOV2 domain of Adiantum phytochrome3, J. Mol. Struct. 735, 259–265. [59] Swartz, T. E., Wenzel, P. J., Corchnoy, S. B., Briggs, W. R., and Bogomolni, R. A. (2002) Vibration spectroscopy reveals light-induced chromophore and protein structural changes in the LOV2 domain of the plant blue-light receptor phototropin 1, Biochemistry 41, 7183–7189. [60] Kandori, H., Belenky, M., and Herzfeld, J. (2002) Vibrational frequency and dipolar orientation

of

the

protonated

Schiff

base

in

bacteriorhodopsin

before

and

after

photoisomerization, Biochemistry 41, 6026–6031. [61] Noguchi, T., Inoue, Y., and Tang, X. S. (1999) Structure of a histidine ligand in the photosynthetic oxygen-evolving complex as studied by light-induced fourier transform infrared difference spectroscopy, Biochemistry 38, 10187–10195.

ACS Paragon Plus Environment

35

Biochemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 37

[62] Noguchi, T., Kurreck, J., Inoue, Y., and Renger, G. (1999) Comparative FTIR analysis of the microenvironment of in cyanide-treated, high pH-treated and iron-depleted photosystem II membrane fragments, Biochemistry 38, 4846–4852. [63] Sasaki, J., Takahashi, H., Furutani, Y., Sineshchekov, O. A., Spudich, J. L., and Kandori, H. (2014) His166 is the Schiff base proton acceptor in attractant phototaxis receptor sensory rhodopsin I, Biochemistry 53, 5923–5929. [64] Iwata, T., Watanabe, A., Iseki, M., Watanabe, M., and Kandori, H. (2011) Strong Donation of the Hydrogen Bond of Tyrosine during Photoactivation of the BLUF Domain, J. Phys. Chem. Lett. 2, 1015–1019. [65] Dürr, H., Salomon, M., and Rüdiger, W. (2005) Chromophore exchange in the LOV2 domain of the plant photoreceptor phototropin1 from oat, Biochemistry 44, 3050–3055. [66] Matsuoka, D., Iwata, T., Zikihara, K., Kandori, H., and Tokutomi, S. (2007) Primary processes during the light-signal transduction of phototropin, Photochem. Photobiol. 83, 122– 130.

ACS Paragon Plus Environment

36

Page 37 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biochemistry

For Table of Contents Use Only

Hydrogen Bonding Environment of the N3–H Group of FMN in the LOV Domains of Phototropins

Tatsuya Iwata,†,# Dai Nozaki,† Atsushi Yamamoto,† Takayuki Koyama,† Yasuzo Nishina,‡ Kiyoshi Shiga,§ Satoru Tokutomi,|| Masashi Unno,⊥ and Hideki Kandori*,†

ACS Paragon Plus Environment

37