Categories
Articles

Spectroscopic and functional characterization of Lampyris turkestanicus luciferase: a comparative study

Original Paper

Pdf

file on Synergy OPEN

omments

Acta Biochim Biophys

Sin 2008, 40: 365-374

doi:10.1111/j.1745-7270.2008.00411.x

Spectroscopic and functional

characterization of Lampyris turkestanicus luciferase: a comparative

study

Mojtaba Mortazavi1,

Saman Hosseinkhani1*, Khosro Khajeh1,

Bijan Ranjbar2, and A. Rahman Emamzadeh1

1 Department of Biochemistry, Faculty of Basic Sciences,

Tarbiat Modares University, Tehran P. O. Box 14115-175, Iran

2 Department of Biophysics, Faculty of Basic

Sciences, Tarbiat Modares University, Tehran P. O. Box 14115-175, Iran

Received: November

26, 2007      

Accepted: March 14,

2008

This work was

supported by a grant from the Ministry of Mines and Industries (Iran) and the

Research Council of Tarbiat Modares University (TMU 85-04-13)

Abbreviations: ANS,

8-anilino-1-naphthalenesulfonic acid; CD, circular dichroism; FT, Fourier transform;

IR, infrared; NTA, nitrilotriacetic acid; UV, ultraviolet

*Corresponding

author: Tel, 98-21-8800-4407; Fax, 98-21-8800-9730; E-mail,

[email protected]

Functional

expression and spectroscopic analysis of luciferases from Lampyris turkestanicus

and Photinus pyralis were carried out. cDNA encoding L. turkestanicus

luciferase was isolated by reverse transcription-polymerase chain reaction,

cloned, and functionally expressed in Escherichia coli. The

luciferases were purified to homogeneity using Ni-nitrilotriacetic acid

Sepharose, and kinetic properties of luciferase from L. turkestanicus

were compared with that from P. pyralis. Amino acid differences

in its primary structures in relation to P. pyralis luciferase

brought about changes in the kinetic properties of the enzyme as evidenced by

substantial lowering of Km for ATP, increased light decay time, and

decreased thermostability. Luciferase from L. turkestanicus was used to

carry out Michaelis-Menten kinetics with a Km of 95.5 mM for ATP and

20 mM for

luciferin. Maximum activity was recorded at pH 8.5, so it might be a suitable

reporter for microbial screening at alkaline pH. Tryptophan fluorescence for P.

pyralis luciferase was higher than L. turkestanicus luciferase.

Substitution of some residues in L. turkestanicus luciferase

appears to change the kinetic properties by inducing a substantial tertiary

structural change, without a large effect on secondary structural elements, as

revealed by intrinsic and extrinsic fluorescence, Fourier transform infrared

spectroscopy, and near-ultraviolet circular dichroism spectra.

Keywords       firefly luciferase; enzyme activity; Lampyris

turkestanicus; emission spectrum; decay rate

Many living organisms are capable of producing visible light, a phenomenon

generally known as bioluminescence. The best-known bioluminescent organism is

the North American firefly, Photinus pyralis, with flash-type

luminescence reaction [1]. The enzyme responsible for the light emission by

firefly is called luciferase, a monomeric enzyme of 62 kDa [2]. Firefly

luciferase (EC 1.13.12.7) catalyzes the oxidation of substrate luciferin in the

presence of ATP, Mg2+, and molecular oxygen

[2,3]. The product, oxyluciferin, is generated in an excited state, then decays

to the ground state with the emission of light. The color of the light emitted

from luminous beetles ranges from green (540 nm) to red (620 nm), and is solely

determined by the active site residues [4,5]. The luciferases have similar

spectra to the color in situ. Each luminous beetle emits a distinctive

flashing pattern that is recognized by the opposite sex in the same species

[6]. The flash may last a few milliseconds, as with the adult Photinus,

or may be a glow lasting several hours, as in the glow worm Lampyris

[7].

The enzyme converts chemical energy, efficiently, into light with a

quantum yield of 0.88 [8]. The sensitivity and convenience of the firefly

luciferase assay has created considerable interest in luciferase-based

biosensors, with a detection limit in the femtomole range. Light production of

firefly luciferase is one of the most sensitive analytical tools in the

ultrasensitive detection of ATP for measuring microbial contamination [9],

genetic reporter assays in molecular biology [10], detection of phosphatase

activity [11], use in DNA sequencing [12], and as a tool for monitoring in

vivo protein folding and chaperonin activity [13].

Assays based on luciferase are preferred due to its high

sensitivity, rapidity, and non-invasiveness. However, several factors limit

further application and development of this technology, including a low

turnover number, high Km for the substrate

ATP, and low stability [14]. For the practical use of luciferase like other

enzymes, both genetic engineering methods such as DNA shuffling and looking for

natural variants are important. Since the first cloning of P. pyralis

luciferase, the luciferase genes have been isolated from several species of

fireflies [15]. The cDNA encoding a glow worm luciferase from lantern mRNA of Lampyris

turkestanicus was cloned and functionally expressed in Escherichia coli

[16].The aim of this work was to characterize the luciferase from L.

turkestanicus and compare its properties with a flash-emitter luciferase

from P. pyralis. Both enzymes were purified to homogeneity; their

enzymatic and structural properties were compared. A further objective was to

determine whether the amino acid sequence or enzyme properties could account

for the characteristics of the glow worm’s glow, in contrast with the flashing

pattern of luciferase from P. pyralis.

Materials and Methods

Chemicals

Mg-ATP and D-luciferin (sodium salt), ATP, and isopropyl-bD-thiogalactopyranoside

were purchased from Sigma (Poole, UK). Ni-nitrilotriacetic acid (NTA) Sepharose

was from Novagen (Madison, USA).

Plasmids and strains

The cDNA of L. turkestanicus luciferase was previously

cloned in pQE30 vector [16]. pET-16b (P. pyralis) was kindly provided by

Prof. Laurence Tisi (University of Cambridge, Cambridge, UK). Both

luciferase-containing vectors (pQE30-luc and pET-16b-luc) were transferred into

E. coli strains XL1-Blue and BL21, respectively. The pQE30-luc

and pET-16b-luc vectors used the T5 and T7 promoters, respectively, and encoded

six histidine residues on the amino terminus of the protein in the multiple

cloning sites that can be used for immobilized metal affinity chromatography.

Transformed colonies by pQE30-luc and pET-16b-uc were screened by X-ray films. Bacteria

containing the pQE30-luc and pET16b-luc were grown overnight at 37 ?C on

Luria-Bertani agar plates containing ampicillin. The master plates were sprayed

with 200 ml solution containing 0.2 mM D-luciferin (in 0.1 M Tris acetate, pH

7.8) for 5 min and exposed to X-ray film for 4 h. After film development, the

positive colonies (bioluminescent) were identified (they also could be observed

by eye after dark-adaptation).

Expression and purification of

recombinant luciferase

A 5 ml culture [2XYT (yeast extract, 10 g; tryptone, 16 g; NaCl, 5

g/L of medium) plus 50 mg/ml ampicillin] was inoculated with a single colony from a freshly

streaked plate of XL1-Blue. The culture was incubated at 37 ?C for 68 h then used to

inoculate a 200 ml fresh culture (2XYT plus 50 mg/ml ampicillin). The 200

ml culture was grown for approximately 5 h at 37 ?C (OD600=0.7). The temperature was then lowered to 22 ?C and the expression

was induced with 1 mM isopropyl-bD-thiogalactopyranoside and 1 mM

lactose. Following overnight induction, the cells were harvested by

centrifugation (12,000 g for 15 min at 4 ?C). The cell pellets were

washed with 10 ml Tris buffer (50 mM, pH 7.8) and repelleted. Cells were

resuspended in 15 ml lysis buffer (10 mM imidazole, 50 mM NaH2PO4, 300 mM NaCl, pH 7.8)

containing 1 mM phenylmethy­lsulfonyl fluoride (Sigma). The cells were lysed by

ultrasonication. Sonication was carried out seven times with 10 s bursts with

the cells on ice to prevent excessive heating of the lysate. The cell lysate

was clarified by centrifugation (18,000 g for 25 min at 4 ?C). The clear

lysate was applied to the Ni-NTA Sepharose column then washed with 20 mM

imidazole in 50 mM NaH2PO4 and 300 mM NaCl (pH 7.8). The histidine-tagged L. turkestanicus

luciferase was eluted with 120250 mM imidazole, desalted using dialysis, and placed into storage

buffer (20% glycerol). The desalted L. turkestanicus luciferase

was concentrated with a Centriprep-10 concentrator (Amicon, Beverly, USA). The

purity of the L. turkestanicus luciferase was more than 95% as

estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. This

process was also carried out to purify P. pyralis luciferase.

Luciferase assay

Luciferase assay

The activity levels were measured using a Sirius Single Tube

Luminometer (Berthold Detection Systems, Pforzheim, Germany) by integration of

total light emitted in 10 s. After the addition of 100 ml standard solution [1 mM

ATP, 0.5 mM D-luciferin, and 10 mM MgSO4 in 50 mM Tris-HCl buffer (pH 7.8)] to 100 ml luciferase-containing

extracts or purified luciferase at 25 ?C, total light units were measured in 10

s.

Optimum temperature

Thermal sensitivity of the luciferases from L.

turkestanicus and P. pyralis were compared. This was carried out

by incubating each luciferase at 10 ?C45 ?C in 50 mM Tricine-NaOH buffer (pH 7.8)

containing 5% glycerol and 10 mM b-mercaptoethanol. Activity was measured at

different temperatures.

Optimum pH

The pH optimums of the luciferases were determined and compared by

preparation of a mix buffer (100 mM glycine, 100 mM succinic acid, and 50 mM

morpholinopropane sulfonic acid). The pH level of the mix buffer was set at 0.5

increments from 5.0 to 12.5. The reaction was initiated by injecting 15 ml diluted luciferase

into 170 ml mix buffer at each pH, then 15 ml substrate was immediately

added and initial activity was measured.

Thermal stability

Thermal stability of the luciferases from L. turkestanicus

and P. pyralis were compared. This was carried out by incubating

enzyme solutions in Eppendorf tubes in a circulating water bath at different

temperatures (20 ?C45 ?C) for 5 min [in 50 mM Tricine-NaOH buffer (pH 7.8) containing

5% glycerol and 10 mM b-mercaptoethanol]. After heating at different temperatures, the

samples were cooled in ice water and remaining activities were assayed

immediately at 25 ?C.

Determination of kinetic

parameters

Luciferase activity was assayed by measuring the peak light emission

with a tube luminometer. The reaction was initiated by injecting 100 ml assay reagent

into 100 ml diluted luciferase. The assay reagent for estimation of Km and Vmax for ATP contained 10 mM MgSO4 and 1 mM luciferin plus 0.1, 0.2, 0.4, 0.8, 1.6, 2.5, 3.5, or 4.5

mM (final concentration) ATP in 50 mM Tricine-NaOH buffer (pH 7.8). The assay

reagent for luciferin Km and Vmax estimation contained 10 mM MgSO4 and 2 mM ATP plus 0.015, 0.025, 0.05, 0.1, 0.2, 0.5, 1.0, or 2.0 mM

luciferin in 50 mM Tricine-NaOH buffer (pH 7.8). The peak light emissions were

taken as a measure of the initial velocity and expressed as relative light

units per second. The data were collected and shown by Michaelis-Menten plots.

The Km and Vmax values for ATP and luciferin were calculated from Lineweaver-Burk

plots. Kinetic experiments were carried out three times and were reproducible

within 5%.

Decay rate

The time-course of the light emitted was measured with a

luminometer. The assay was carried out by addition of 50 ml substrate [1

mM luciferin, 50 mM Tris buffer, 2 mM ATP, and 10 mM MgSO4 (pH 7.8)] into a 50 ml purified luciferase solution from P.

pyralis or L. turkestanicus. The estimated mixing time was

1.0 s, then data was acquired.

Measurement of bioluminescence emission spectra

Bioluminescence emission spectra of the luciferases from P.

pyralis and L. turkestanicus were measured with an LS 50B

luminescence spectrophotometer (PerkinElmer Optoelectronics, Fremont, USA). A

volume of 2 ml substrate mixture consisting of 1 mM luciferin, 50 mM Tris

buffer (pH 7.8), 2 mM ATP, and 10 mM MgSO4 was added to 50 ml purified luciferase solution in a quartz cell. Data were collected

over the wavelength range 400700 nm. The spectra were automatically corrected for the

photosensitivity of the equipment.

Fluorescence measurements

Tryptophan fluorescence was measured on the LS 50B luminescence

spectrophotometer. The excitation wavelength was set at 295 nm and the emission

spectra were obtained. Extrinsic fluorescence studies were carried out as

previously described [17] using 8-anilino-1-naphthalenesulfonic acid (ANS) as a

fluorescence probe. Measurements were taken on the same spectrofluorometer as

used for intrinsic fluorescence studies. All experiments were carried out at 25

?C with ANS and protein concentrations of 30 mM and 1 mM, respectively,

in 50 mM phosphate buffer. An excitation wavelength of 350 nm was used.

Infrared (IR) spectroscopy

Fourier transforms (FT)-IR spectra were recorded on a Nexus 870

FT-IR spectrophotometer (Thermo Fisher Scientific, Waltham, USA) with a

deuterated triglycine sulfate (DTGS) detector. The samples (1 mM) were placed

in a cell. Typically, 1 mM protein mixture was injected into the cell and a spectrum

recorded. The cell was then drained and the protein deposited on the crystal

was dried. Second-derivative spectra were calculated by a method reported

previously for bacterial luciferase [18].

Circular dichroism (CD)

measurements

CD spectra were recorded on a Jasco J-715 spectropolarimeter (Tokyo,

Japan) using solutions with protein concentrations of approximately 0.2 and 1.5

mg/ml for far- and near-ultraviolet (UV) regions, respectively. Results are expressed

as molar ellipticity, [q] (deg?cm2?mol1), based on a mean amino

acid residue weight (MRW) of 115 for luciferase. The molar ellipticity was

determined as [q]=(q100MRW)/(cl), where c is the protein concentration in

mg/ml, l is the light path length in centimeters, and q is the measured

ellipticity in degrees at a wavelength q. The instrument was

calibrated with (+)-10-camphorsulfonic acid, assuming [q]291=7820 deg?cm2?mol1, and with Jasco standard

non-hydroscopic ammonium (+)-10-camphorsulphonate, assuming [q]290.5=7910 deg?cm2?mol1. The data were smoothed

using the fast FT noise reduction routine that allows enhancement of most noisy

spectra without distorting their peak shapes.

Calculation of essential

residues pKa

The data of crystal structure of Japanese firefly (Luciola

cruciata) luciferase were downloaded from Protein Data Bank (2D1Q; http://www.pdb.org/pdb/home/home.do).

The crystal structure of Japanese thermostable firefly luciferase was used as a

template to make a suitable model for L. turkestanicus and P.

pyralis. We used the software package MacroDox (Northrup, Cookeville, USA)

for our analysis [19]. The modeling data were then used to calculate pKa

of critical residues involved in enzyme catalysis or substrate binding.

Results

Purification of the

polyhistidine-tagged luciferases

The cDNA encoding both luciferases containing a polyhistidine tag (6?his) at the amino terminus of the protein were used to express

luciferases. Cells were lysed and the clarified cytoplasmic extracts applied on

a column. After washing with low imidazole (20 mM), unbound and weakly bound

proteins were removed. The polyhistidine-tagged luciferase was finally eluted

from the column by increasing the imidazole concentration gradient up to 250

mM. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis and

Coomassie Brilliant Blue staining of the eluted fractions showed that the

polyhistidine-tagged enzymes were efficiently bound to the column and that the

corresponding fractions contained highly purified protein (more than 95%).

Determination of protein contents in each fraction was carried out using the

Bradford method. These results clearly showed that the enzymes were

expressed at relatively high levels (approximately 10 mg per liter culture),

and were efficiently purified, and that the enzymes had retained their

biological activity [22].

Decay rate

The time-course of the light emitted by both luciferases was also

measured for glow worm and firefly luciferases (Fig. 1). The decay in

light emission was different for each enzyme under the same conditions.

However, the rate of luminescence decay for P. pyralis firefly

luciferase was faster, suggesting lower stability of the intermediate structure

of the luciferase reaction (Table 1).

Bioluminescence emission

spectra

The color of the light emitted by the L. turkestanicus

luciferase was green as that of the P. pyralis luciferase. Fig.

2 compares the bioluminescence emission spectra of these luciferases in the

presence of luciferin and ATP at pH 7.8. Emission spectra of luciferases from L.

turkes­tanicus and P. pyralis showed a high similarity

without changing the lmax (Fig. 2) and only a minor broadening of the spectrum for P.

pyralis luciferase was observed.

Optimum temperature of

luciferase

The thermosensitivity of the P. pyralis and L.

turkestanicus luciferases was compared. The activity of both luciferases

reached a maximum at 25 ?C and started to decrease at higher temperatures (Fig.

3). When the temperature increased above the thermal transition temperature

(approximately 45 ?C), both luciferases lost almost all of their activity.

Thermostability of luciferase

The thermostability of the P. pyralis and L.

turkestanicus luciferases was also compared. The remaining activity of

luciferase from P. pyralis reached its maximum at 25 ?C, whereas

the activity of luciferase from L. turkestanicus started to decrease

at this temperature (Fig. 4). When the temperature increased above the

thermal transition temperature (approximately 45 ?C), both luciferases lost

almost all of their activity. The thermostability of these enzymes, however,

showed that luciferase from L. turkestanicus has lower

temperature stability than luciferase from P. pyralis.

pH optimums

The pH optimum of these luciferases was determined and

compared.  The pH optimum for luciferase

from P. pyralis was 8.0, whereas the pH optimum for luciferase

from L. turkestanicus showed a maximum at 8.5 (Fig. 5).

Therefore, comparative study of these enzymes confirmed a higher optimum pH for

L. turkestanicus when compared to P. pyralis.

Kinetic properties of purified

luciferases

Table 1 compares some properties of L.

turkestanicus and P. pyralis luciferases. The bioluminescent

activity of luciferase is directly proportional to the concentration of ATP

present in the reaction mixture [9,20]. Therefore, extensive washing procedures

are necessary to eliminate ATP and other components originating from the host organism [21]. For L.

turkestanicus luciferase the Km for luciferin was 20 mM; for P.

pyralis luciferase the Km for [chshou1] luciferin was 30 mM. However, for L. turkestanicus luciferase the Km for ATP was 95.5 mM, whereas for P. pyralis luciferase it was 140 mM. The Vmax values for different concentrations of ATP were similar for both

luciferases. It should be noted that subcloning of the luciferases in

the same vector (pET-28a) did not change their kinetic properties.

Spectra of luciferases

CD spectra of both luciferases obtained in Tricine buffer (50 mM, pH

7.8) are shown in Fig. 6. As indicated in Fig. 6(A), the far-UV

CD spectra of luciferase showed a small decrease in negative ellipticity at 208

nm and 222 nm in L. turkestanicus luciferase. The near-UV CD

spectra indicated that the defined tertiary structure of the L.

turkestanicus luciferase was decreased compared to P. pyralis

luciferase [Fig. 6(B)].

Fluorescence measurements

Intrinsic and ANS fluorescence spectroscopy were used to

characterize the microenvironments of Trp and also hydrophobic clusters of

luciferases. As indicated in Fig. 7, an increase of fluorescence

intensity was observed for P. pyralis luciferase. Loss of

hydrophobic patches on the L. turkestanicus protein surface was

confirmed by fluorescence measurements using ANS (Fig. 8). ANS

fluorescence was clearly enhanced on interaction with native luciferase, as

observed earlier [22].

IR spectra

Fig. 9 compares the second-derivative

spectrum of luciferases at pH 7.8. In 1H2O, the amide I band

showed a maximum at 1652 cm1 for both luciferases. The bands at 1658, 1666, and 1672 have been

assigned to a-structures, b-turns, and b-sheets, respectively. Changes in luciferase structure seemed to

cause a small modification in the amount of a-structures and b-turns. The band

at 1658 cm1 had been previously

assigned to the C=O stretching vibration of the side chain amide groups. The

most obvious difference was the intensity of the band around 1658 cm1, which was visibly a

shift to longer wave numbers in L. turkestanicus luciferase.

Discussion

The firefly luciferase gene, luc, is frequently used as a

reporter of genetic function, or for producing of recombinant luciferase [23].

To simplify the purification process, luciferases with six histidine residues

on their N-terminus were used. The luciferase bound to Ni-NTA Sepharose through

the strong chelating interaction between the His tag and Ni2+ directly from the cell lysate. A high purity recombinant luciferase

with a yield of 95% was achieved. The result reported earlier showed that the

luciferase responsible for the light emitting reaction in L.

turkestanicus, although three residues shorter, has 85% sequence

similarity with that of the firefly P. pyralis [16]. Both enzymes

have a similar C-terminus with the same key residues.

Kinetic differences

According to its kinetic properties, L. turkestanicus luciferase

could be considered a suitable indicator for ATP detection. In fact, the Km of L. turkestanicus luciferase for ATP was lower than

P. pyralis luciferase, whereas the Km towards luciferin was almost identical (Table 1). Using

ProSite [24], putative interacting residues with AMP in both luciferases have

been identified as 195-IMNSSGSTGLPK-206 [16]. The crystal structure of P.

pyralis luciferase shows that invariant residues Arg218, His245, Phe247,

Ala348, and Lys529 appear to interact with luciferin in luciferase binding site

[25]. Furthermore, the residues Asn197, Ser199, Thr343, Tyr340, Ala317, and

Gly339 in the P. pyralis luciferase sequence are believed to

interact with ATP [26]. Using homology modeling, the same residues have been

found to interact with ATP in L. turkestanicus luciferase (data

not shown). Therefore, it could be suggested that the higher affinity of L.

turkestanicus luciferase for ATP might be due to specific orientation of

these residues in the active site of the enzyme or conformational changes. The time-course of the light decay for both luciferases was also

compared (Fig. 1, Table 1). The rate of light decay for L.

turkestanicus, similar to another glow worm luciferase, was slower than P.

pyralis [27]. However, P. pyralis firefly luciferase produced

a reproducibly faster decay, suggesting faster decomposition of the

intermediate state of the luciferase reaction or more sensitivity to product

inhibition [27].

Similar light emission spectra of the L. turkestanicus

and P. pyralis luciferases (Fig. 2) also suggests

similarities in critical residues involved in the formation of substrate

intermediate structures. One of the main reasons for differences in the

luciferase bioluminescence color is the property of the emitter

microenvironment localized in the enzyme active site [28]. The variety of

bioluminescence color is also attributed to the luciferase structure [29]. However,

it seems the emitter environments of both luciferases are similar.Another intriguing result is that the shape of the bioluminescence

spectrum for L. turkestanicus luciferase under acidic conditions

was slightly changed [30]. In contrast to most firefly (Lampyridae)

luciferases, luciferase from L. turkestanicus showed a minor

shift in its emission peak under this condition, similar to that of

pH-insensitive luciferases (click beetles and railroad worms). Although firefly

luciferases have high amino acid similarity, multiple sequence alignment of L.

turkestanicus luciferase with other firefly luciferases revealed certain

amino acid changes in L. turkestanicus that are essentially

sufficient for it to have a pH-independent luminescence spectra profile. Among

them is Phe268, a conserved residue in firefly luciferases, which has been

substituted by Cys in L. turkestanicus [31].

The optimum temperature of both luciferases was similar (Fig. 3).

However, the increased temperature sensitivity of the glow worm luciferase (Fig.

4) might make it a slightly better reporter for the study of gene control

elements for high turnover mRNAs and unstable proteins in live cells [32]. The

precise mechanism responsible for the low thermostability of luciferase from L.

turkestanicus is not fully understood. The lower thermostability of L.

turkestanicus luciferase compared to P. pyralis luciferase

was similar to another glow worm luciferase, as reported earlier [27]. Another property that should be considered in the application of L.

turkestanicus luciferase is its optimum pH. As indicated in Fig. 5,

its optimum pH is 0.5 higher than P. pyralis luciferase.

Calculation of the pKa of essential residues showed that most of the

critical residues involved in binding of substrates or catalysis had a higher pKa

in L. turkestanicus compared to P. pyralis (Table

2). Therefore, it could be suggested that the higher optimum pH of L.

turkestanicus luciferase could arise from the higher pKa of some

critical-reported residues.

Structural differences Characterization of L. turkestanicus luciferases

containing an N-terminal His-tag by spectroscopic devices revealed differences

with P. pyralis luciferase. It has been shown that the His-tag

has very little effect on the structure, activity, or stability of firefly

luciferases [33]. Differences in primary structure of L.

turkestanicus compared to P. pyralis luciferase have changed

the protein conformation, as confirmed by intrinsic fluorescence. For example, L.

turkestanicus luciferase has only one Trp at position 419, whereas P.

pyralis has two Trp residues. W419 can be used as a suitable reporter in

the study of L. turkestanicus luciferase conformational changes

under different conditions. Therefore, as shown in Fig. 7, differences in

primary structure and the presence of an additional Trp caused a clear increase

in intrinsic fluorescence of P. pyralis luciferase, suggesting

that Trp(s) are located in a more hydrophobic environment. A minor red shift in

the emission spectra was also seen, suggesting that other factors might be

involved in fluorescence emission. A similar result has been observed in

changes of fluorescence intensity in native and a mutant form of bacterial

luciferase [18,34]. The hydropathy profiles of these luciferases were similar,

however, there were some regions that showed major differences. In particular,

the regions from residue 140 to 200 and from 300 to 340, which indicates the

additional hydrophobic residues in P. pyralis luciferase, showed

major differences in hydrophobic pattern when compared to L.

turkestanicus luciferase. The presence of some substitution in the L.

turkestanicus luciferase primary structure is associated with decreased

surface hydrophobicity compared to P. pyralis luciferase, as was

also confirmed by ANS fluorescence spectroscopy (Fig. 8). Such a feature

could be potentially involved with considerable conformational changes among

these luciferases. It has been reported that such changes could affect the

bioluminescence colors and other kinetic properties of luciferases [35]. The

far-UV and near-UV CD spectra of both forms were taken, and they reflected

substantial changes in the tertiary structure of the protein with small changes

in its secondary structure (Fig. 6). The type of changes observed in the

CD and IR spectra of luciferases suggests conformational differences.In conclusion, results presented in this study show that L.

turkestanicus luciferase, similar to P. pyralis luciferase,

was successfully expressed in sufficient amounts in E. coli.

Moreover, the kinetic properties of L. turkestanicus, including

higher affinity to ATP, show it can be used as a suitable reporter in molecular

biology.

Acknowledgements

We thank

Ms. Z. Fazl-Zarandi (Department of Biochemistry, Tarbiat Modares University,

Tehran, Iran) for her helpful cooperation.

References

 1   Buck J. Synchronous rhythmic flashing of

fireflies II. Q Rev Bio 1988, 63: 265289

 2   DeLuca M, McElroy WD. Purification and

properties of firefly luciferase. Methods Enzymol 1978, 57: 315

 3   Seliger HH, McElroy WD. Quantum yield in the

oxidation of firefly luciferin. Biochem Biophys Res Commun 1959, 1: 2124

 4   Seliger HH, McElroy WD. The colors of firefly

bioluminescence: enzyme configuration and species specificity. Proc Natl Acad

Sci USA 1964, 52: 7581

 5   Viviani VR, Bechara EJH. Biophysical and

biochemical aspects of phengodid (railroadworm) bioluminescence. Photochem

Photobiol 1993, 58: 615622

 6   Lloyd JE. Insect bioluminescence. In: Herring

PJ, ed. Bioluminescence in Action. London: Academic 1978, 241272

 7   Newport G. On the natural history of the

glowworm (Lampyris noctiluca). J Proc Linn Soc 1957, 1: 4071

 8   Seliger HH, McElroy WD. Spectral emission and

quantum yield of firefly bioluminescence. Arch Biochem Biophys 1960, 88: 136145

 9   Lundin A. Use of firefly luciferase in

ATP-related assays of biomass, enzymes, and metabolites. Methods Enzymol 2000,

305: 346370

10  Gould SJ, Subramani S. Firefly luciferase as a

tool in molecular and cell biology. Anal Biochem 1988, 175: 513

11  Miska W, Geiger R. A new type of ultra

sensitive bioluminogenic enzyme substrates. I. Enzyme substrates with

D-luciferin as leaving group. Biol Chem Hoppe Seyler 1988, 369: 40711

12  Ronaghi M, Karamohamed S, Petterson B, Uhlen M,

Nyren P. Real-time DNA sequencing using detection of pyrophosphate release.

Anal Biochem 1996, 242: 8489 

13  Frydman J, Nimmesgern E, Erdjumentbromage H,

Wall JS, Tempst P, Hartl FU. Function in protein folding of TRiC, a cytosolic

ring complex containing TCP-1 and structurally related subunits. EMBO J 1992,

11: 47674778

14  Kajiyama N, Nakano E. Thermostabilization of

firefly luciferase by a single amino acid substitution at position 217.

Biochemistry 1993, 32: 1379513799

15  Branchini BR, Southworth TL, DeAngelis JP,

Roda A, Michelini E. Luciferase from the Italian firefly Luciola italica:

Molecular cloning and expression. Comp Biochem Physiol B Biochem Mol Biol

2006, 145: 159167

16  Alipour BS, Hosseinkhani S, Nikkhah M,

Naderi-Manesh H, Chaichi MJ, Osaloo SK. Molecular cloning, sequence analysis

and expression of a cDNA encoding the luciferase from the glow-worm Lampyris

turkestanicus. Biochem Biophys Res Commun 2004, 325: 215222

17  Hosseinkhani S, Ranjbar B, Naderi-Manesh H,

Nemat-Gorgani M. Chemical modification of glucose oxidase: possible formation

of molten globule-like intermediate structure. FEBS Lett 2004, 561: 213216

18  Riahi-Madvar A, Hosseinkhani S, Khajeh K,

Ranjbar B, Asoodeh BA. Implication of a critical residue (Glu175) in structure

and function of bacterial luciferase. FEBS Lett 2005, 579: 47014706

19  Northrup SH. MacroDox v.2.0.2: Software for

the Prediction of Macromolecular Interaction. Cookeville: Tennessee

Technological University, 1995


20  Rajgopal S, Vijayalakshmi MA. Firefly luciferase:

purification and immobilization. Enzyme Microb Technol 1984, 6: 482490

21  Michel P, Torkkeli T, Karp M, Oker-Blom C.

Expression and purification of polyhistidine-tagged firefly luciferase in

insect cells – a potential alternative for process scale-up. J Biotechnol 2001,

85: 4956

22 Yousefi-Nejad M, Hosseinkhani S, Khajeh K,

Ranjbar B. Expression, purification and immobilization of firefly luciferase on

alkyl-substituted Sepharose 4B. Enzyme Microb Technol 2007, 40: 740746

23  Sherf BA, Wood KV. Firefly luciferase

engineered for improved genetic reporting. Promega Notes 1994, 49: 1421

24  Gattiker A, Gasteiger E, Bairoch A.

ScanProsite: a reference implementation of a PROSITE scanning tool. Appl

Bioinformatics 2002, 1: 107108

25  Viviani VR, Bechara EJH, Ohmiya Y. Cloning,

sequence analysis, and expression of active Phrixothrix railroad-worms

luciferases: relationship between bioluminescence spectra and primary

structure. Biochemistry 1999, 38: 82718279

26  Hirokawa K, Kajiyama N, Murakami S. Improved

practical usefulness of firefly luciferase by gene chimerization and random

mutagenesis. Biochim Biophys Acta 2002, 1597: 271279

27  Sala-Newby GB, Thomson CM, Campbell AK.

Sequence and biochemical similarities between the luciferase of the glow-worm Lampyris

noctiluca and the firefly Photinus pyralis. Biochem J 1996, 313: 761767

28  Campbell AK. Chemiluminescence: Principles and

Applications in Biology and Medicine. Chichester: Ellis Horwood 1988

29  Viviani VR, Uchida A, Suenaga N, Ryufuku M,

Ohmiya Y. Thr226 is a key residue for bioluminescence spectra determination in

beetle luciferases. Biochem Biophys Res Commun 2001, 280: 12861291

30  Tafreshi NK, Hosseinkhani S, Sadeghizadeh M,

Sadeghi M, Ranjbar B, Naderi-Manesh H. The influence of insertion of a critical

residue (Arg356) in structure and bioluminescence spectra of firefly

luciferase. J Biol Chem 2007, 282: 86418647

31  Viviani VR. The origin, diversity, and

structure function relationships of insect luciferases. Cell Mol Life Sci 2002,

59: 18331850

32  Thompson JF, Hayes LS, Lloyd DB. Modulation of

firefly luciferase stability and impact on studies of gene regulation. Gene

1991 103: 171177

33  Wang WQ, Xu Q, Shan YF, Xu GJ. Probing local

conformational changes during equilibrium unfolding of firefly luciferase:

Fluorescence and circular dichroism studies of single tryptophan mutants.

Biochem Biophys Res Commun 2001, 282: 2833

34  Hosseinkhani S, Szittner R, Meighen E. Random

mutagenesis of bacterial luciferase: Critical role of Glu175 in control of

luminescence decay. Biochem J 2005, 385: 575580

35  Viviani VR, Bechara EJH. Bioluminescence of

Brazilian fireflies (coleoptera: Lampyridae): spectral distribution and

pH effect on luciferase-elicited colors. Comparison with elaterid and phengodid

luciferases. Photochem Photobiol 1995, 62: 490495