23.12.2012 Views

Aim of the thesis - VU-DARE Home - Vrije Universiteit Amsterdam

Aim of the thesis - VU-DARE Home - Vrije Universiteit Amsterdam

Aim of the thesis - VU-DARE Home - Vrije Universiteit Amsterdam

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

VRIJE UNIVERSITEIT<br />

Molecular aspects <strong>of</strong> <strong>the</strong> histamine H1 receptor<br />

ACADEMISCH PROEFSCHRIFT<br />

ter verkrijging van de graad Doctor aan<br />

de <strong>Vrije</strong> <strong>Universiteit</strong> <strong>Amsterdam</strong>,<br />

op gezag van de rector magnificus<br />

pr<strong>of</strong>.dr. L.M. Bouter,<br />

in het openbaar te verdedigen<br />

ten overstaan van de promotiecommissie<br />

van de faculteit der Exacte Wetenschappen<br />

op dinsdag 1 juni 2010 om 11.45 uur<br />

in de aula van de universiteit,<br />

De Boelelaan 1105<br />

door<br />

Kamonchanok Sansuk<br />

geboren te Bangkok, Thailand


promotor: pr<strong>of</strong>.dr. R. Leurs<br />

copromotor: dr. R.A. Bakker


To my parents,<br />

who brought me into this world,<br />

and shaped me into who I am,


ISBN: 9789461080394<br />

Cover:Tibetan Thangka “Purification <strong>of</strong> <strong>the</strong> mind”<br />

Photoshop by Rodrigo Reijers<br />

Printed at: Gildeprint B.V., Enschede, <strong>the</strong> Ne<strong>the</strong>rlands


Contents<br />

Chapter 1 Introduction to <strong>the</strong> histamine H1 receptor<br />

AIM <strong>of</strong> <strong>the</strong> <strong>the</strong>sis<br />

Chapter 2 Expression and purification <strong>of</strong> GPCRs<br />

Chapter 3 Expression <strong>of</strong> <strong>the</strong> histamine H1 and H4<br />

receptor by in vitro cell free expression<br />

system and <strong>the</strong>ir mass spectrometric<br />

analysis<br />

Chapter 4 Mass spectrometric analysis <strong>of</strong> <strong>the</strong> H1R:<br />

glycosylation and binding <strong>of</strong> an<br />

irreversible H1R antagonist<br />

Chapter 5 A structural insight into <strong>the</strong> reorientation<br />

<strong>of</strong> transmembrane domains 3 and 5<br />

during H1R activation<br />

Chapter 6 The role <strong>of</strong> isoleucine at position 6.45 in<br />

<strong>the</strong> H1R<br />

Chapter 7 General discussion<br />

Summary<br />

Samenvatting<br />

References<br />

Acknowledgements<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]<br />

p7-24<br />

p25-26<br />

p27-50<br />

p51-70<br />

p71-84<br />

p85-104<br />

p105-116<br />

p117-122<br />

p123-126<br />

p127-130<br />

p131-146<br />

P147-150


CHAPTER<br />

Introduction to <strong>the</strong> histamine H1 receptor<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Chapter 1<br />

8 | P a g e<br />

Introduction to <strong>the</strong> H1R<br />

The histamine H1 receptor (H1R) belongs to <strong>the</strong> membrane protein family <strong>of</strong> G protein-coupled<br />

receptors (GPCRs). GPCRs or also known as seven-transmembrane domain (7TM) receptors<br />

represent <strong>the</strong> largest class <strong>of</strong> membrane protein in <strong>the</strong> human genome (Jacoby, Bouhelal et al. 2006;<br />

Schlyer and Horuk 2006; Lefkowitz 2007; Lagerstrom and Schioth 2008). All members <strong>of</strong> this<br />

membrane protein family, including <strong>the</strong> H1R, share a common membrane topology: an extracellular<br />

N-terminus, an intracellular C-terminus and seven transmembrane (TM) helices, which are<br />

connected by three intra-cellular loops (ICLs) and three extracellular loops (ECLs) (Figure 1)<br />

(Jacoby, Bouhelal et al. 2006; Schlyer and Horuk 2006; Lefkowitz 2007; Lagerstrom and Schioth<br />

2008).<br />

Figure 1. GPCR structure overview. This figure is obtained from Nygaard (Nygaard, Frimurer et<br />

al. 2009).<br />

Serpentine model <strong>of</strong> a GPCR with <strong>the</strong> seven transmembrane (TM) helical segments (TM-I to TM-VII). The<br />

highly conserved disulfide bridge from <strong>the</strong> end <strong>of</strong> TM-III to <strong>the</strong> middle <strong>of</strong> extracellular loop 2 (ECL-2). The<br />

model also show a helix VIII at <strong>the</strong> C-terminus. Conserved residues are indicated with orange (conserved) and<br />

red (highly conserved) symbols. The three main residues are R3.50 <strong>of</strong> TM-III, W6.48 <strong>of</strong> TM-VI and Y7.53 <strong>of</strong><br />

TM-VII. Their presumed interaction partners are indicated by dotted lines and arrows (Nygaard, Frimurer et<br />

al. 2009).


Introduction to <strong>the</strong> histamine H1 receptor<br />

The H1R is expressed ubiquitously in lung smooth muscle, vascular endo<strong>the</strong>lium, gastro-intestinal<br />

tract, heart, and brain. Histamine, which is an endogenous ligand <strong>of</strong> <strong>the</strong> H1R, is derived from<br />

histidine and stored in basophils, mast cells, lymphocytes, neurons, and gastric enterochromaffin-<br />

like cells, (Barger and Dale 1910; Thurmond, Gelfand et al. 2008). Through <strong>the</strong> H1R , histamine<br />

exerts many actions related to allergic inflammation (e.g. local reddening, wheal and flare <strong>of</strong> skin,<br />

nasal congestion, and bronchoconstriction) (Thurmond, Gelfand et al. 2008) and CNS effects (e.g.<br />

reduction <strong>of</strong> locomotor activity, exploratory behavior, and cognitive functions) (Inoue, Yanai et al.<br />

1996).<br />

Signaling pathway <strong>of</strong> GPCRs and <strong>the</strong> H1R<br />

Figure 2. GPCR signaling. This figure is modified from Siehler (Siehler 2008).<br />

The GPCRs coupled to Gs proteins stimulate adenylyl cyclase (AC) causing an increase in intracellular cyclic<br />

adenosine mono-phosphate (cAMP) levels, whereas <strong>the</strong> GPCRs coupled to Gi proteins inhibit <strong>the</strong> formation <strong>of</strong><br />

cAMP. cAMP regulates various cellular function, through <strong>the</strong> activation <strong>of</strong> protein kinases such as protein<br />

kinase A (PKA). The GPCRs coupled to Gq protein activate phospholipase C-β (PLC-β), resulting in <strong>the</strong><br />

hydrolysis <strong>of</strong> phosphatidylinositol-4, 5-biphosphate (PIP2) to inositol-1,4,5-triphosphate (IP3) and<br />

diacylglycerol (DAG). DAG increase <strong>the</strong> activity <strong>of</strong> PKC, and IP3 triggers <strong>the</strong> release <strong>of</strong> Ca 2+ from intracellular<br />

stores. The GPCRs coupled to G12 activates PLC-ε and <strong>the</strong> small GTPase Rho.<br />

9 | P a g e


Chapter 1<br />

10 | P a g e<br />

H1R agonists<br />

H1R antagonists<br />

Figure 3. Chemical structures <strong>of</strong> various H1R ligands


Introduction to <strong>the</strong> histamine H1 receptor<br />

Originally, GPCR signaling is thought to be dependent on association with a heterotrimeric G<br />

protein. Interaction <strong>of</strong> an activated GPCR with a G protein catalyzes <strong>the</strong> exchange <strong>of</strong> guanosine<br />

diphosphate (GDP) by guanosine triphosphate (GTP), enabling Gα- and βγ-subunits to interact with<br />

a variety <strong>of</strong> downstream effectors (Siehler 2008). Signaling <strong>of</strong> <strong>the</strong> Gα subunit is dependent on <strong>the</strong><br />

Gα is<strong>of</strong>orm as shown in Figure 2. All <strong>the</strong> four members <strong>of</strong> <strong>the</strong> Gαq family (Gαq, Gα11, Gα14, and<br />

Gα15/16) stimulate phospholipase C-β (PLC-β) enzymes (Berridge 1993; Rhee 2001; Offermanns<br />

2003). The βγ-subunits also activate PLC-β, although with significantly lower affinities than <strong>the</strong> Gαq<br />

(Rhee 2001). PLC-β <strong>the</strong>n catalyzes <strong>the</strong> hydrolysis <strong>of</strong> phosphatidylinositol diphosphate (PIP2) to<br />

form inositol triphosphate (IP3) and diacylglycerol (DAG) (Berridge 1993; Rhee 2001). Activating <strong>of</strong><br />

protein kinase C (PKC) by DAG modulates <strong>the</strong> activity <strong>of</strong> many cellular substrates including <strong>the</strong><br />

transcription factors like NF-κB (Berridge 1993; Marinissen and Gutkind 2001). IP3 on <strong>the</strong> o<strong>the</strong>r<br />

hand, binds to IP3 receptors at <strong>the</strong> membrane <strong>of</strong> endoplasmic reticulum and causes <strong>the</strong> release <strong>of</strong><br />

Ca 2+ into <strong>the</strong> cytosol (Siehler 2008). The more recently described Gα12 stimulates phospholipase C-ε<br />

(PLC-ε) causing increases <strong>of</strong> intracellular IP3 and DAG (Kelley, Kaproth-Joslin et al. 2006). In<br />

addition to <strong>the</strong> activation <strong>of</strong> PLC-ε, Gα12 can also activate PKC through <strong>the</strong> small GTPase Rho (Sah,<br />

Seasholtz et al. 2000; Siehler 2008). Gαs and Gαi proteins mostly modulate adenylyl cyclase (AC)<br />

activity (Birnbaumer and Birnbaumer 1995; Offermanns 2003). The activation <strong>of</strong> AC results in <strong>the</strong><br />

formation <strong>of</strong> cAMP, which regulates o<strong>the</strong>r proteins such as protein kinase A (PKA) (Siehler 2008).<br />

However, recent studies (Daaka, Luttrell et al. 1997; Vanhauwe, Thomas et al. 2002; Marinissen,<br />

Servitja et al. 2003; Defea 2008) suggest that GPCR signaling is far more diverse than originally<br />

thought. It is now generally accepted that GPCRs can couple to multiple G proteins (Daaka, Luttrell<br />

et al. 1997). For example, <strong>the</strong> β2AR can couple to both Gαs and Gαi (Daaka, Luttrell et al. 1997) and<br />

<strong>the</strong> protease-activated receptor-1 (PAR-1) can couple to Gαi/o, Gαq and Gα12/13 (Vanhauwe, Thomas<br />

et al. 2002; Marinissen, Servitja et al. 2003). These complicated couplings <strong>of</strong> GPCRs demonstrate <strong>the</strong><br />

importance <strong>of</strong> evaluating multiple signaling pathways as one assay for <strong>the</strong> quantification <strong>of</strong><br />

receptor activation may not be sufficient. An increase in G protein coupling in response to agonist<br />

binding can still be evaluated by quantifying membrane-associated GTPase activity. However, over<br />

<strong>the</strong> last few years it has been shown that GPCRs can even signal without coupling to a G protein (see<br />

review by (Defea 2008)). This G protein independent signaling is mediated through β-arrestins and<br />

G-protein coupled receptor kinases (GRK) and can result in <strong>the</strong> activation <strong>of</strong> cytoplasmic effectors<br />

such as Src, Ras and extracellular signal-regulated kinase ERK (Lefkowitz and Shenoy 2005; Reiter<br />

and Lefkowitz 2006; Defea 2008).<br />

11 | P a g e


Chapter 1<br />

The H1R mediates its action through <strong>the</strong> Gαq/11 protein (Leopoldt, Harteneck et al. 1997). As<br />

described previously, <strong>the</strong> Gαq/11 protein activates PLC, which <strong>the</strong>n hydrolyses PIP2 into IP3 and DAG.<br />

Subsequently, DAG activates PKC, which in turn can phosphorylate o<strong>the</strong>r proteins, including <strong>the</strong><br />

H1R, resulting in receptor desensitization. IP3 causes <strong>the</strong> release <strong>of</strong> Ca 2+ from <strong>the</strong> intracellular store.<br />

The elevation <strong>of</strong> intracellular Ca 2+ concentration can explain a variety <strong>of</strong> pharmacological<br />

responses; for example <strong>the</strong> induction <strong>of</strong> smooth muscle cell contraction, <strong>the</strong> activation <strong>of</strong> nitric<br />

oxide synthase (Satoh and Inui 1984), and <strong>the</strong> stimulation <strong>of</strong> phospholipase A2 (PLA2) (Murayama,<br />

Kajiyama et al. 1990). Besides <strong>the</strong> Gαq/11 protein, <strong>the</strong> H1R can signal via <strong>the</strong> Gαi/o protein (Seifert,<br />

Grunbaum et al. 1994) and is also reported to activate <strong>the</strong> transcription factor NF-κB (Leurs,<br />

Church et al. 2002) ei<strong>the</strong>r via βγ subunits in constitutively activate H1R or via both Gαq/11 and βγ<br />

subunits (Bakker, Schoonus et al. 2001).<br />

Figure 4. Interaction <strong>of</strong> retinal and a rotamer toggle switch <strong>of</strong> rhodopsin. This figure is<br />

modified from Bhattacharya (Bhattacharya, Hall et al. 2008).<br />

(A) Chemical structure <strong>of</strong> 11-cis-retinal and all-trans-retinal<br />

(B) In <strong>the</strong> inactive state, <strong>the</strong> aromatic moiety <strong>of</strong> tryptophan W265 (W6.48) is parallel to <strong>the</strong> lipid bilayer<br />

(Left). The indole nitrogen <strong>of</strong> <strong>the</strong> W6.48 also forms a water mediated hydrogen bond with N302 (N7.49) <strong>of</strong> <strong>the</strong><br />

conserved NPxxY motif <strong>of</strong> TM7. In <strong>the</strong> active state, <strong>the</strong> hydrogen bond with TM7 is disrupted when <strong>the</strong><br />

flipping <strong>of</strong> <strong>the</strong> W6.48 is triggered by <strong>the</strong> isomerization <strong>of</strong> 11-cis-retinal to all-trans-retinal.<br />

12 | P a g e


Introduction to <strong>the</strong> histamine H1 receptor<br />

Figure 5. Crystal structures <strong>of</strong> Rhodopsin, β2AR and A2A adenosine receptor. This figure is<br />

obtained from Hanson (Hanson and Stevens 2009).<br />

The ECL-2 region (Top): The extracellular domain in rhodopsin serves to occlude <strong>the</strong> retinal binding pocket,<br />

whereas in both <strong>the</strong> β2AR and A2A adenosine receptors <strong>the</strong> extracellular domain is highly constrained and<br />

held away from <strong>the</strong> ligand binding pocket.<br />

The binding pocket region (middle): The retinal binding pocket relies on hydrophobic interactions in addition<br />

to a covalent linkage with lysine residue in TM5. The tail <strong>of</strong> carazolol, <strong>the</strong> β2AR ligand, forms polar<br />

interactions with a conserved glutamate on TM3 and asparagine on TM7, whereas <strong>the</strong> head <strong>of</strong> carazolol forms<br />

polar interactions with serine residues on TM5. The adenosine ligand ZM241385 forms polar interactions<br />

with asparagine residues on TM6 and a glutamate on ECL-2<br />

The intracellular interactions (Bottom): bovine rhodopsin is <strong>the</strong> only receptor with an intact ionic lock<br />

interaction between <strong>the</strong> E/DRY motif and glutamate on TM6. All <strong>the</strong> o<strong>the</strong>r two structures have a disrupted<br />

ionic lock. Thee E/DRY motif <strong>of</strong> β2AR and A2A adenosine receptor interacts with ICL-2 through a polar<br />

interaction between <strong>the</strong> aspartate residue and ei<strong>the</strong>r a serine or tyrosine residue on ICL-2.<br />

13 | P a g e


Chapter 1<br />

Fur<strong>the</strong>rmore, GPCRs, including <strong>the</strong> H1R, can also activate <strong>the</strong>ir G protein in <strong>the</strong> absence <strong>of</strong> <strong>the</strong>ir<br />

ligands. This basal activity is so called “constitutive activity” (Jacoby, Bouhelal et al. 2006; Kobilka<br />

and Deupi 2007; Lefkowitz 2007). Many GPCRs, which undergo spontaneous mutations and posses<br />

constitutive activity, can actually cause a variety <strong>of</strong> human illnesses (Spiegel 1998; Jacoby, Bouhelal<br />

et al. 2006; Lefkowitz 2007). For example, <strong>the</strong> hypercalcemia and skeletal dysplasia in Jansen’s<br />

disease is <strong>the</strong> result <strong>of</strong> a constitutively active parathyroid hormone (PTH) receptor (Schipani, Kruse<br />

et al. 1995), <strong>the</strong> familial male precocious puberty is caused by a single point mutation in <strong>the</strong><br />

luteinizing hormone (LH) receptor (Shenker, Laue et al. 1993) and <strong>the</strong> congenital night blindness is<br />

due to mutations in rhodopsin (Dryja, Berson et al. 1993). The discovery <strong>of</strong> constitutive activity has<br />

also led to a reclassification <strong>of</strong> ligands. Full agonists are capable <strong>of</strong> maximal receptor stimulation.<br />

Partial agonists are unable to give full activity. Neutral antagonists have no effect on signaling<br />

activity (but can prevent o<strong>the</strong>r ligands from binding to a receptor). Full and partial inverse agonists<br />

can reduce <strong>the</strong> level <strong>of</strong> basal or constitutive activity (Bond, Leff et al. 1995; Leurs, Church et al.<br />

2002; Strange 2002; Kobilka and Deupi 2007).<br />

14 | P a g e<br />

The H1R ligands<br />

Since <strong>the</strong> discovery <strong>of</strong> <strong>the</strong> H1R, numerous histamine derivatives have been syn<strong>the</strong>sized to enhance<br />

agonist potency and selectivity. At first, <strong>the</strong> imidazole ring <strong>of</strong> histamine (Figure 3) was replaced by<br />

o<strong>the</strong>r heterocyclic rings. The most prominent compounds found in this class are 2-(thiazol-2-yl)-<br />

ethylamine and 2-pyridylethylamine (Donaldson and Hill 1986). Ano<strong>the</strong>r strategy which focused on<br />

shortening or elongation <strong>of</strong> <strong>the</strong> ethylamine side-chain, however, resulted in loss <strong>of</strong> potency (Zingel,<br />

Leschke et al. 1995). The introduction <strong>of</strong> different substitutes at <strong>the</strong> C-2 position <strong>of</strong> <strong>the</strong> imidazole<br />

ring resulted in selective H1R agonists, such as 2-(3-bromophenyl)histamine and 2-(3-<br />

trifluoromethyl)-phenyl-histamine (Leschke, Elz et al. 1995). Interestingly, <strong>the</strong> addition <strong>of</strong> a<br />

diphenylalkyl group, which is typically found in H1R antagonists, also lead to a more selective and<br />

potent H1R agonist, called histaprodifen (Elz, Kramer et al. 2000). Ano<strong>the</strong>r interesting ligand is 8R-<br />

lisuride, which is a potent H1R agonist that is derived from ergot alkaloids. This compound is,<br />

however, not selective since it is also known to bind dopamine and serotonin receptors. All <strong>the</strong>se<br />

H1R agonists may serve as valuable tools for <strong>the</strong> investigation <strong>of</strong> H1R mediated physiological and<br />

pathological functions (Bakker, Weiner et al. 1994). The first generation antihistamines generally<br />

contain <strong>the</strong> ethylamine moiety linked to two aromatic rings (Figure 3) such as mepyramine,<br />

chlorpheniramine, hydroxyzine and diphenhydramine. However, fusion <strong>of</strong> <strong>the</strong> tricyclic rings is also


Introduction to <strong>the</strong> histamine H1 receptor<br />

tolerated (e.g. doxepin and ketotifen). The first generation antihistamines were effective through<br />

<strong>the</strong>ir strong histamine inhibiting activity. However, <strong>the</strong>y also induce sedation and psychomotor<br />

retardation among allergic patients, due to <strong>the</strong>ir ability to cross <strong>the</strong> blood brain barriers and bind to<br />

muscarinic receptors or H1R in <strong>the</strong> central nervous system (Thurmond, Gelfand et al. 2008). The<br />

second generation H1R antagonists were <strong>the</strong>refore developed with strongly reduced side effects.<br />

Several <strong>of</strong> <strong>the</strong> second generation antihistamines such as cetirizine, loratadine and fex<strong>of</strong>enadine are<br />

in use for <strong>the</strong> treatment <strong>of</strong> allergic symptoms.<br />

Inactive-state GPCRs<br />

Structural knowledge <strong>of</strong> GPCRs and <strong>the</strong> H1R<br />

The first insights into <strong>the</strong> structure <strong>of</strong> GPCRs came from <strong>the</strong> crystal <strong>of</strong> bovine rhodopsin in year<br />

2000 (See fur<strong>the</strong>r details on protein expression at chapter 2) (Palczewski, Kumasaka et al. 2000).<br />

Rhodopsin is well suited for structural studies due to its high expression level in nature (bovine<br />

retina) and high stability compared to o<strong>the</strong>r GPCRs (Nygaard, Frimurer et al. 2009). Rhodopsin<br />

consists <strong>of</strong> <strong>the</strong> 7TM protein “opsin” and <strong>the</strong> chromophore “11-cis-retinal” (Figure 4A), which is<br />

covalently bound to <strong>the</strong> lysine residue K7.43 in TM7. The β-ionone ring (at <strong>the</strong> o<strong>the</strong>r end <strong>of</strong> <strong>the</strong> 11-<br />

cis-retinal) is buried deeply in a hydrophobic pocket and has a direct interaction with <strong>the</strong><br />

tryptophan W6.48 (Palczewski, Kumasaka et al. 2000; Nygaard, Frimurer et al. 2009) from <strong>the</strong><br />

CWxP motif in TM6. This tryptophan is a part <strong>of</strong> important intra-molecular interaction known as a<br />

“rotamer toggle switch”, which consists <strong>of</strong> residues F6.52, C6.47 and W6.48 (Nygaard, Frimurer et<br />

al. 2009). Upon exposure to light, 11-cis-retinal isomerizs to all-trans-retinal, which induces a series<br />

<strong>of</strong> conformational changes in <strong>the</strong> receptor, including a switching <strong>of</strong> <strong>the</strong> rotamer conformation <strong>of</strong> <strong>the</strong><br />

W6.48 into its active form (Figure 4B) (Bhattacharya, Hall et al. 2008; Nygaard, Frimurer et al.<br />

2009). With <strong>the</strong> information obtained from <strong>the</strong> crystal structure <strong>of</strong> rhodopsin (Palczewski,<br />

Kumasaka et al. 2000), earlier models <strong>of</strong> GPCRs based on <strong>the</strong> crystal structure <strong>of</strong><br />

bacteriorhodopsins (Luecke, Schobert et al. 1999) have been re-evaluated. In analogy to <strong>the</strong><br />

lipophillic binding pocket <strong>of</strong> 11-cis- retinal in <strong>the</strong> crystal structure <strong>of</strong> rhodopsin (Palczewski,<br />

Kumasaka et al. 2000), <strong>the</strong> main ligand binding site <strong>of</strong> small molecule hormones and non-peptides<br />

is thought to locate within <strong>the</strong> central crevice <strong>of</strong> <strong>the</strong> 7TM domains (Smit, Vischer et al. 2007;<br />

Bhattacharya, Hall et al. 2008; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009).<br />

Fur<strong>the</strong>rmore, <strong>the</strong> crystal structure <strong>of</strong> rhodopsin also revealed <strong>the</strong> proline kinks <strong>of</strong> TM1, 4 and 7<br />

15 | P a g e


Chapter 1<br />

(Menon, Han et al. 2001), <strong>the</strong> bending <strong>of</strong> TM2 by a pair <strong>of</strong> G2.56 and G2.57 (Stenkamp, Filipek et al.<br />

2002), <strong>the</strong> short eighth helix (Palczewski, Kumasaka et al. 2000) and <strong>the</strong> β-sheet at <strong>the</strong> ECL2 which<br />

forms a lid toge<strong>the</strong>r with N-terminus on top <strong>of</strong> <strong>the</strong> chromophore binding (Bourne and Meng 2000).<br />

Figure 6. The superimposition <strong>of</strong> selected GPCR structures. This figure is obtained from Topiol<br />

(Topiol and Sabio 2009).<br />

(A) View from <strong>the</strong> toggle switch region (W6.48) <strong>of</strong> three GPCRs: <strong>the</strong> bovine rhodopsin with retinal (in<br />

brown), <strong>the</strong> human β2AR with carazolol (in pink) and <strong>the</strong> human A2A adenosine receptor with ZM241385 (in<br />

green). The residues corresponding to phenylalanine (F290) in <strong>the</strong> human β2AR is alanine (A) in <strong>the</strong> bovine<br />

rhodopsin and histidine (H) in <strong>the</strong> human A2A adenosine receptor.<br />

(B) View from <strong>the</strong> intracellular side <strong>of</strong> bovine rhodopsin and opsin with bound GαCT. TM6 (VI) <strong>of</strong> opsin<br />

(yellow) is tilted outward by 6-7 Å. The glutamic residue E247 decouples from <strong>the</strong> ionic lock interaction with<br />

R135 (R3.50) from <strong>the</strong> highly conserved DRY motif <strong>of</strong> TM3, allowing TM6 to move out and away from TM3<br />

(III) and swing over to TM5 (V). This E247, <strong>the</strong>n forms a salt bridge with K231 <strong>of</strong> TM5, helping TM6 to come<br />

closer to TM5. The Y306 (Y7.53) from <strong>the</strong> NPxxY motif <strong>of</strong> TM7 is rotated into <strong>the</strong> helix core, blocking <strong>the</strong><br />

inward return <strong>of</strong> TM6.<br />

The rhodopsin crystals stayed <strong>the</strong> only GPCR structure resolved to an atomic level for ano<strong>the</strong>r<br />

seven years until <strong>the</strong> crystal structures <strong>of</strong> human β2 adrenergic receptor (β2AR) (Cherezov,<br />

Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007), <strong>the</strong> turkey β1 adrenergic receptor (β1AR)<br />

(Warne, Serrano-Vega et al. 2008) and human A2A adenosine receptors (Jaakola, Griffith et al. 2008)<br />

were resolved with <strong>the</strong> use <strong>of</strong> a variety <strong>of</strong> different protein-modification and engineering<br />

approaches (see fur<strong>the</strong>r details on protein expression at chapter 2). It has been shown that <strong>the</strong> four<br />

16 | P a g e


Introduction to <strong>the</strong> histamine H1 receptor<br />

inactive state GPCR structures are overall in close agreement. The most significant structural<br />

differences lie in <strong>the</strong> extracellular, intracellular and ligand-binding regions. It is known from <strong>the</strong><br />

rhodopsin crystals that <strong>the</strong> ECL2 region appears to form a β-sheet structure on top <strong>of</strong> <strong>the</strong><br />

chromophore binding cavity (Figure 5, ECL-2) (Palczewski, Kumasaka et al. 2000) and maybe act<br />

as a lid shielding 11-cis-retinal from hydrolysis (Audet and Bouvier 2008; Hanson, Cherezov et al.<br />

2008; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). However, in <strong>the</strong> β2-AR and β1AR X-<br />

ray structures, <strong>the</strong> ECL2 loop is a short α-helix with cysteine bridges (Cherezov, Rosenbaum et al.<br />

2007; Rasmussen, Choi et al. 2007; Warne, Serrano-Vega et al. 2008). It has been hypo<strong>the</strong>sized that<br />

<strong>the</strong> structural features <strong>of</strong> ECL2 region in β2AR and β1AR are more accessible to <strong>the</strong> extracellular<br />

environment and allow <strong>the</strong> diffusion <strong>of</strong> ligands in and out <strong>of</strong> <strong>the</strong> binding pocket (Hanson, Cherezov<br />

et al. 2008; Hanson and Stevens 2009; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009).<br />

The ECL2 region <strong>of</strong> <strong>the</strong> A2A receptor, however, resembles nei<strong>the</strong>r <strong>the</strong> β-sheet <strong>of</strong> rhodopsin nor <strong>the</strong><br />

short α-helix <strong>of</strong> adrenergic receptors. In <strong>the</strong> region, <strong>the</strong> receptor adopts a random coil conformation<br />

and has multiple cysteine bridges, in which three <strong>of</strong> <strong>the</strong>m are linked to <strong>the</strong> ECL1 and one is linked to<br />

<strong>the</strong> ECL3 (Jaakola, Griffith et al. 2008). This structural feature results in an open structure at <strong>the</strong><br />

extracellular end that can accommodate <strong>the</strong> entry <strong>of</strong> ligands (Hanson, Cherezov et al. 2008; Hanson<br />

and Stevens 2009; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). Additional<br />

structures will be needed to confirm whe<strong>the</strong>r <strong>the</strong> open binding pocket is a general feature <strong>of</strong> <strong>the</strong><br />

ligand-activated GPCRs.<br />

Based on <strong>the</strong> crystal structures <strong>of</strong> rhodopsin, adrenergic and adenosine receptors; ligand binding is<br />

mediated by polar and hydrophobic contact residues from TM3, 5, 6 and 7 (Cherezov, Rosenbaum<br />

et al. 2007; Rasmussen, Choi et al. 2007; Jaakola, Griffith et al. 2008; Warne, Serrano-Vega et al.<br />

2008). In rhodopsin (Figure 5, Binding pocket), 11-cis-retinal is located deep in <strong>the</strong> pocket with<br />

its β-ionone ring reaching down between <strong>the</strong> indole side-chain <strong>of</strong> W6.48 (Palczewski, Kumasaka et<br />

al. 2000). The adrenergic antagonists bind to <strong>the</strong>ir receptors similar to that <strong>of</strong> 11-cis retinal, but<br />

<strong>the</strong>y do not reach as deep into <strong>the</strong> pocket as retinal does. With regard to <strong>the</strong> toggle switch,<br />

carazolol, a β2AR antagonist, interacts indirectly with <strong>the</strong> W6.48 via <strong>the</strong> phenylalanine F290<br />

(Figure 6A) and force <strong>the</strong> tryptophan residue to be in its inactive conformation (Topiol and Sabio<br />

2009). In contrast to <strong>the</strong> adrenergic and rhodopsin structures, ZM241385, an adenosine A2A<br />

receptor antagonist, binds vertically to <strong>the</strong> bilayer plane <strong>of</strong> <strong>the</strong> A2A receptor and fur<strong>the</strong>r away from<br />

<strong>the</strong> W6.48. However, <strong>the</strong> receptor is still held in <strong>the</strong> inactive conformation via recruitment <strong>of</strong> H250<br />

as an intermediate residue (Figure 6A) (Topiol and Sabio 2009). The observed differences between<br />

17 | P a g e


Chapter 1<br />

<strong>the</strong> crystal structures <strong>of</strong> adrenergic and adenosine receptors suggest that <strong>the</strong> ligand-activated<br />

GPCRs may have a wide variety <strong>of</strong> ligand binding modes (Hanson and Stevens 2009; Nygaard,<br />

Frimurer et al. 2009; Topiol and Sabio 2009).<br />

At <strong>the</strong> cytoplasmic surface, a major difference between <strong>the</strong> ligand-activated GPCRs and rhodopsin<br />

lies in ei<strong>the</strong>r <strong>the</strong> presence or absence <strong>of</strong> <strong>the</strong> polar interaction networks (also known as “ionic lock”)<br />

between <strong>the</strong> highly conserved E/DRY motif on TM3 and a glutamate residue on TM6 (Hanson and<br />

Stevens 2009; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). This ionic lock is<br />

conserved among all <strong>the</strong> rhodopsin crystal structures and is thought to stabilize <strong>the</strong> inactive-state<br />

conformation (Ballesteros, Jensen et al. 2001; Kobilka and Deupi 2007; Audet and Bouvier 2008;<br />

Hanson and Stevens 2009; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). The ionic<br />

lock has been observed in <strong>the</strong> inactive state rhodopsin crystals (Palczewski, Kumasaka et al. 2000),<br />

but not in all <strong>the</strong> ligand-activated GPCRs structures (Cherezov, Rosenbaum et al. 2007; Rasmussen,<br />

Choi et al. 2007; Jaakola, Griffith et al. 2008; Warne, Serrano-Vega et al. 2008). This can be<br />

interpreted that ei<strong>the</strong>r this interaction is so weak that it is broken during crystallization or it does<br />

not exist in <strong>the</strong> ligand-bound states (Hanson and Stevens 2009; Nygaard, Frimurer et al. 2009;<br />

Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). Interestingly, among <strong>the</strong> notable<br />

differences found between <strong>the</strong> β1AR and β2AR structures are <strong>the</strong> residues in <strong>the</strong> ionic lock region. In<br />

<strong>the</strong> β1AR crystal, a short α-helix is formed in <strong>the</strong> ICL2 region, and <strong>the</strong>reby accommodating a<br />

hydrogen bond between Y149 on ICL2 and <strong>the</strong> aspartate <strong>of</strong> <strong>the</strong> E/DRY motif in TM3 (Warne,<br />

Serrano-Vega et al. 2008). This structural feature is also observed in <strong>the</strong> A2A receptor crystal<br />

(Jaakola, Griffith et al. 2008), but not in <strong>the</strong> β2AR and rhodopsin crystals (Palczewski, Kumasaka et<br />

al. 2000; Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007) (Figure 5, bottom). The<br />

absence <strong>of</strong> a short α-helix and <strong>the</strong> hydrogen bond network in <strong>the</strong> β2AR structure can be used to<br />

explain <strong>the</strong> higher constitutive activity <strong>of</strong> <strong>the</strong> β2AR compared to that <strong>of</strong> <strong>the</strong> β1AR (Nygaard,<br />

Frimurer et al. 2009; Rosenbaum, Rasmussen et al. 2009).<br />

Active-state GPCRs<br />

The activation <strong>of</strong> <strong>the</strong> rhodopsin family is thought to involve <strong>the</strong> disruption <strong>of</strong> intra-molecular<br />

interactions that stabilize <strong>the</strong>ir inactive conformations (Smit, Vischer et al. 2007; Bhattacharya, Hall<br />

et al. 2008; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). Based on homology<br />

modeling (Bhattacharya, Hall et al. 2008) and biophysical experiments with several GPCRs<br />

18 | P a g e


Introduction to <strong>the</strong> histamine H1 receptor<br />

(Dunham and Farrens 1999; Stru<strong>the</strong>rs, Yu et al. 2000; Altenbach, Klein-Seetharaman et al. 2001;<br />

Huang, Li et al. 2001; Huang, Visiers et al. 2002; Shi, Liapakis et al. 2002), such as site-directed<br />

mutagenesis (Huang, Li et al. 2001; Huang, Visiers et al. 2002), cysteine accessibility (Shi, Liapakis<br />

et al. 2002), disulfide cross-linking (Stru<strong>the</strong>rs, Yu et al. 2000), spin labeling (Altenbach, Klein-<br />

Seetharaman et al. 2001) and circular dichroism (Dunham and Farrens 1999), <strong>the</strong> cytoplasmic ends<br />

<strong>of</strong> TM3 and TM6 are proposed to move away from one ano<strong>the</strong>r during <strong>the</strong> activation process due to<br />

<strong>the</strong> breaking <strong>of</strong> <strong>the</strong> ionic lock between <strong>the</strong> highly conserved E/DRY motif on TM3 and a glutamate<br />

residue on TM6. The R3.50 is also restrained in <strong>the</strong> inactive conformation by a hydrophobic<br />

arginine cage formed by conserved amino acids at position 3.46 (L: 15%, V: 18%, I: 58%, M: 15%)<br />

and 6.37 (L: 37%, V: 24%, I: 20%, M: 5%) (Smit, Vischer et al. 2007). The breaking <strong>of</strong> <strong>the</strong> ionic lock<br />

<strong>of</strong>ten results in constitutive activation <strong>of</strong> several GPCRs such as rhodopsin (Kim, Altenbach et al.<br />

1997), H2R (Alewijnse, Timmerman et al. 2000), α1B adrenergic receptor (Scheer, Fanelli et al.<br />

1996), luteinizing hormone receptor (Laue, Chan et al. 1995) and thyrotropin receptor<br />

(Ringkananont, Van Durme et al. 2006). Ano<strong>the</strong>r important intra-molecular interaction known as a<br />

“rotamer toggle switch”, which consists <strong>of</strong> residues F6.52, C6.47 and W6.48 has also been proposed<br />

to be involved in <strong>the</strong> activation process <strong>of</strong> <strong>the</strong> rhodopsin family (Smit, Vischer et al. 2007;<br />

Bhattacharya, Hall et al. 2008; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio 2009). In <strong>the</strong><br />

inactive conformation, <strong>the</strong> plane <strong>of</strong> <strong>the</strong> aromatic moiety <strong>of</strong> W6.48 is parallel to <strong>the</strong> lipid bilayer,<br />

while in <strong>the</strong> predicted active state, it is perpendicular to <strong>the</strong> lipid bilayer (Figure 4B)<br />

(Bhattacharya, Hall et al. 2008). This rotamer flipping hypo<strong>the</strong>sis is supported by <strong>the</strong> information<br />

from <strong>the</strong> crystal structure <strong>of</strong> Meta I (<strong>the</strong> first stable intermediate <strong>of</strong> photo-activated rhodopsin) that<br />

showed <strong>the</strong> toggling <strong>of</strong> <strong>the</strong> tryptophan rotamer (Ruprecht, Mielke et al. 2004). The toggling <strong>of</strong><br />

W6.48 is thought to modulate <strong>the</strong> bend angle around <strong>the</strong> highly conserved proline P6.50 on TM6<br />

and leading to a movement <strong>of</strong> <strong>the</strong> cytoplasmic end <strong>of</strong> TM6 during <strong>the</strong> activation process (Shi,<br />

Liapakis et al. 2002; Vilardaga 2006). In <strong>the</strong> inactive state, <strong>the</strong> highly conserved N7.49 <strong>of</strong> <strong>the</strong> NPxxY<br />

motif in TM7 is restrained toward TM6 via water molecule in many rhodopsin family GPCRs (Smit,<br />

Vischer et al. 2007). This asparagine residue is located in a hydrophobic cage formed by conserved<br />

amino acids at position 2.46 (L: 91%) and 6.40 (L: 14%, V: 42%, I: 28%, M: 5%) (Smit, Vischer et al.<br />

2007). It was proposed that upon receptor activation, N7.49 adopts <strong>the</strong> trans (t) conformation to<br />

interact with D2.50 <strong>of</strong> <strong>the</strong> (N/S)LxxxD motif in TM2 (Urizar, Claeysen et al. 2005), and could force<br />

<strong>the</strong> relocation <strong>of</strong> R3.50 from <strong>the</strong> ionic lock (Bakker, Jongejan et al. 2008). Mutations <strong>of</strong> asparagine at<br />

position 7.49 or removal <strong>of</strong> <strong>the</strong> bulky and hydrophobic cage affects <strong>the</strong> constitutive activity <strong>of</strong><br />

several GPCRs such as rhodopsin (Han, Lin et al. 1996; Madabushi, Gross et al. 2004), thyroid<br />

19 | P a g e


Chapter 1<br />

stimulating hormone receptor (Urizar, Claeysen et al. 2005), serotonin 5HT2A receptor (Shapiro,<br />

Kristiansen et al. 2002) and H1R (Bakker, Jongejan et al. 2008). Shortly, it is suggested that a<br />

mechanism for receptor activation may involve a separation <strong>of</strong> <strong>the</strong> TM3 and TM6, toge<strong>the</strong>r with a<br />

twist in TM6 which pulls <strong>the</strong> ICL3 into <strong>the</strong> cell and uncovers residues related to G-protein coupling<br />

(Rovati, Capra et al. 2007; Audet and Bouvier 2008). This hypo<strong>the</strong>sis is fur<strong>the</strong>r supported by <strong>the</strong><br />

current crystal structure <strong>of</strong> opsin (Park, Scheerer et al. 2008). Opsin is <strong>the</strong> retinal-free<br />

photoreceptor generated after <strong>the</strong> photo-activation and Schiff base hydrolysis <strong>of</strong> rhodopsin. There<br />

are several subtle changes in <strong>the</strong> conformations <strong>of</strong> binding pocket residues, relative to rhodopsin:<br />

<strong>the</strong> side chain <strong>of</strong> W6.48 moves into space previously occupied by 11-cis retinal, <strong>the</strong> interaction<br />

between K7.43 and E3.28 is broken, <strong>the</strong> cytoplasmic end <strong>of</strong> TM6 is shifted outwards from <strong>the</strong> centre<br />

<strong>of</strong> <strong>the</strong> TM bundle and became closer to TM5, <strong>the</strong> ionic lock is broken and Y7.53 from <strong>the</strong> NPxxY<br />

motif on TM7 undergoes a conformational change and inserts into space occupied by TM6 in <strong>the</strong><br />

inactive state rhodopsin (Figure 6B). All <strong>the</strong>se changes from inactive rhodopsin to active-state<br />

opsin result in <strong>the</strong> creation <strong>of</strong> a cavity between TM3, TM5 and TM6, in which <strong>the</strong> G protein can bind<br />

(Topiol and Sabio 2009).<br />

Inactive-state H1R<br />

Similar to o<strong>the</strong>r members <strong>of</strong> GPCRs, with <strong>the</strong> exception <strong>of</strong> rhodopsin, adrenergic receptors and<br />

adenosine receptor, structural analysis at atomic level <strong>of</strong> <strong>the</strong> H1R are still limited by low expression<br />

levels in both nature and heterologous expression systems and/or by difficulties in obtaining pure<br />

and homogenous protein for crystallization. To date, <strong>the</strong> crystal structure <strong>of</strong> H1R has not been<br />

solved yet. As a consequence, most <strong>of</strong> <strong>the</strong> structural studies <strong>of</strong> <strong>the</strong> H1R were only possible via<br />

mutational analysis and computational techniques based on homology modeling with available<br />

GPCR crystals. Based on various site-directed mutational analysis and homology modeling studies<br />

(Leurs, Smit et al. 1994; Ohta, Hayashi et al. 1994; Moguilevsky, Varsalona et al. 1995; ter Laak,<br />

Timmerman et al. 1995; Wieland, Laak et al. 1999; Jongejan, Bruysters et al. 2005; Jongejan and<br />

Leurs 2005), <strong>the</strong> endogenous ligand histamine is believed to bind various amino acid residues in<br />

TM3, TM5 and TM6 <strong>of</strong> <strong>the</strong> H1R (Figure 8). The H1R has a well conserved aspartate residue D3.32 in<br />

TM3 that is believed to be one <strong>of</strong> <strong>the</strong> most crucial amino acids for ligand binding among biogenic<br />

amine receptors. The residue D3.32 <strong>of</strong> <strong>the</strong> H1R is thought to form an ionic bond with <strong>the</strong> amino<br />

group <strong>of</strong> histamine (Ohta, Hayashi et al. 1994; Jongejan and Leurs 2005). Several studies fur<strong>the</strong>r<br />

demonstrated that <strong>the</strong> asparagine N5.46 <strong>of</strong> <strong>the</strong> H1R is involved in ligand binding pocket by forming<br />

20 | P a g e


Introduction to <strong>the</strong> histamine H1 receptor<br />

an interaction with <strong>the</strong> distal nitrogen position (N τ ) <strong>of</strong> <strong>the</strong> imidazole ring <strong>of</strong> histamine (Leurs, Smit<br />

et al. 1994; Moguilevsky, Varsalona et al. 1995; Jongejan and Leurs 2005), while <strong>the</strong> lysine residue<br />

K5.39 interacts with <strong>the</strong> ortho nitrogen (N π ) <strong>of</strong> <strong>the</strong> imidazole ring <strong>of</strong> histamine (Leurs, Smit et al.<br />

1995; ter Laak, Timmerman et al. 1995; Gillard, Van Der Perren et al. 2002; Jongejan and Leurs<br />

2005). Mutation <strong>of</strong> residue K5.39 in guinea pig (Leurs, Smit et al. 1995), or in human (Gillard, Van<br />

Der Perren et al. 2002) H1R resulted in a decrease in histamine potency and binding affinity.<br />

Bruysters and co-workers also demonstrated that <strong>the</strong> serine S3.36 plays a part in <strong>the</strong> H1R binding<br />

pocket and acts as a toggle switch in receptor activation mechanism (Jongejan, Bruysters et al.<br />

2005). Fur<strong>the</strong>rmore, <strong>the</strong> residue F6.55 may be important in stabilizing <strong>the</strong> imidazole ring <strong>of</strong><br />

histamine, since mutation <strong>of</strong> F6.55 to alanine in both guinea pig and human resulted in a reduction<br />

<strong>of</strong> histamine potency (Wieland, Laak et al. 1999). In short, amino acid residues, that are involved in<br />

<strong>the</strong> histamine binding pocket consist <strong>of</strong> D3.32, S3.36 <strong>of</strong> TM3; K5.39, N5.46 <strong>of</strong> TM5 and F6.55 <strong>of</strong><br />

TM6. Bruysters and co-workers (Bruysters, Pertz et al. 2004) found that <strong>the</strong> residues D3.32 and<br />

F6.52 are also crucial for <strong>the</strong> binding <strong>of</strong> histaprodifen, a selective H1R agonist that posses <strong>the</strong><br />

structural features <strong>of</strong> both agonist and antagonist. However, <strong>the</strong> binding <strong>of</strong> histaprodifen was not<br />

affected by <strong>the</strong> mutation at residues K5.39 and F6.55. Fur<strong>the</strong>rmore, <strong>the</strong> mutation <strong>of</strong> N5.46 does not<br />

influence <strong>the</strong> affinity <strong>of</strong> histaprodifen, although <strong>the</strong> signaling as measured by reporter gene assay is<br />

greatly impaired. Interestingly, Bruysters and co-workers (Bruysters, Jongejan et al. 2005) also<br />

observed <strong>the</strong> species selectivity <strong>of</strong> histaprodifen upon employing <strong>the</strong> mutant H1R, in which “human”<br />

amino acids were individually replaced by <strong>the</strong>ir guinea pig counterparts. Mutation <strong>of</strong> N2.61 to<br />

serine resulted in a complete reversal <strong>of</strong> human to guinea pig pharmacology. The conserved Y7.43<br />

is also found to be pivotal in <strong>the</strong> mechanism behind <strong>the</strong> species selectivity. Based on <strong>the</strong>se data, it<br />

can be concluded that not all H1R agonists bind in a similar manner. The binding pocket for H1R<br />

agonist may not be limited to only TM3, 5 and 6, but may also include TM2 and 7, as shown in<br />

Figure 8 (Jongejan and Leurs 2005).<br />

For antagonist binding, <strong>the</strong> residue D3.32 also appears to be an important interaction point.<br />

Mutation <strong>of</strong> D3.32 to alanine resulted in a complete loss <strong>of</strong> [ 3H]mepyramine binding affinity (Ohta,<br />

Hayashi et al. 1994; Nonaka, Otaki et al. 1998). However, <strong>the</strong> radioligand [ 3H]KW-4679 was shown<br />

to bind to <strong>the</strong> D3.32A H1R mutant with a reasonable affinity compared to that <strong>of</strong> <strong>the</strong> wild type H1R<br />

(D3.32A: KD = 49 nM and wild type: KD = 2.5 nM), showing that <strong>the</strong> tricyclic H1R antagonists such as<br />

KW-4679 may have different binding pocket than mepyramine (Nonaka, Otaki et al. 1998).<br />

Mutation <strong>of</strong> N5.46 to alanine resulted in an unaltered affinity <strong>of</strong> [ 3H]mepyramine,<br />

21 | P a g e


Chapter 1<br />

chlorpheniramine, cetirizine, and loratadine, suggesting that histamine and its antagonists may<br />

interact with different amino acids in <strong>the</strong> same binding pocket (Leurs, Smit et al. 1994; Ohta,<br />

Hayashi et al. 1994; Moguilevsky, Varsalona et al. 1995).<br />

Figure 8. Binding <strong>of</strong> H1R ligands to <strong>the</strong> H1R. The figure is obtained from Jongejan (Jongejan and<br />

Leurs 2005).<br />

(A) The H1R model was obtained using <strong>the</strong> rhodopsin template. The amino group <strong>of</strong> histamine form an<br />

interaction with D3.32 <strong>of</strong> TM3. The distal nitrogen position (N τ ) <strong>of</strong> <strong>the</strong> imidazole ring <strong>of</strong> histamine forms an<br />

interaction with N5.46 <strong>of</strong> TM 5. The ortho nitrogen (N π ) <strong>of</strong> <strong>the</strong> imidazole ring <strong>of</strong> histamine forms an<br />

interaction with K5.39 <strong>of</strong> TM5.<br />

(B) The classical binding site (light gray) showing residues involved in binding <strong>of</strong> classical H1R agonists:<br />

D3.32, K5.39, T5.42, N5.46, F6.52 and F6.55. The binding pocket <strong>of</strong> histaprodifen (darker grey) extends<br />

from <strong>the</strong> classical pocket to TM2 and TM7.<br />

The introduction <strong>of</strong> a carboxylic moiety in many second generation H1R antagonists showed to limit<br />

brain penetration, thus reducing sedative side effects. Despite <strong>the</strong> beneficial pharmacokinetic<br />

properties, this carboxylic moiety also seems to affect <strong>the</strong> binding affinity. Mutation <strong>of</strong> <strong>the</strong> guinea<br />

pig H1R at <strong>the</strong> positively charged K5.39 to <strong>the</strong> uncharged alanine resulted in strong decreased<br />

affinities <strong>of</strong> acrivastine and levocetirizine, but not meclozine and tripolidine which are <strong>the</strong>ir analogs<br />

that lack <strong>the</strong> carboxy moiety (Wieland, Laak et al. 1999). Interestingly, <strong>the</strong> mutation <strong>of</strong> K5.39 to <strong>the</strong><br />

positively charged arginine resulted in a wild type phenotype again, demonstrating <strong>the</strong> importance<br />

<strong>of</strong> <strong>the</strong> positive charge <strong>of</strong> K5.39 position for an interaction with <strong>the</strong> negatively charged carboxylate<br />

group <strong>of</strong> acrivastine and levocetirizine (Wieland, Laak et al. 1999).<br />

22 | P a g e


Introduction to <strong>the</strong> histamine H1 receptor<br />

Figure 9. Binding <strong>of</strong> H1R antagonists to <strong>the</strong> H1R. The figure is obtained from Wieland (Wieland,<br />

Laak et al. 1999).<br />

(A) The H1R antagonist pharmacophore describing <strong>the</strong> position <strong>of</strong> <strong>the</strong> cis- and trans-aromatic rings.<br />

(B) The H1R antagonist (cyproheptadine) was docked into <strong>the</strong> H1R model based on BR (ter Laak, Timmerman<br />

et al. 1995). A view from <strong>the</strong> extracellular side shows <strong>the</strong> orientation <strong>of</strong> cyproheptadine between <strong>the</strong> TM<br />

domain (yellow) and <strong>the</strong> ionic hydrogen bond interaction with Asp116 (D3.32) in TM III (TM3). The transaromatic<br />

ring <strong>of</strong> cyproheptadine is in <strong>the</strong> proximity <strong>of</strong> Phe433 (F6.52), Phe436 (F6.55) <strong>of</strong> TM VI (TM6), and<br />

<strong>the</strong> cis-aromatic ring is near Trp161 (W6.47) <strong>of</strong> TM IV (TM4).<br />

Using a 3D model <strong>of</strong> <strong>the</strong> H1R receptor based on <strong>the</strong> crystal structure <strong>of</strong> bovine rhodopsin (ter Laak,<br />

Timmerman et al. 1995) and site-directed mutagenesis analysis (Wieland, Laak et al. 1999),<br />

Wieland and co-workers showed that residues W4.67 <strong>of</strong> TM4, F6.52 and F6.55 <strong>of</strong> TM6 are<br />

responsible for <strong>the</strong> interaction with <strong>the</strong> aromatic rings <strong>of</strong> <strong>the</strong> H1R antagonists. The cis-ring <strong>of</strong><br />

cyproheptadine is located within 5 Å <strong>of</strong> a tryptophan W4.67 <strong>of</strong> TM4, whereas <strong>the</strong> trans-aromatic<br />

ring <strong>of</strong> cyproheptadine is close to two phenylalanine residues F6.52 and F6.55 <strong>of</strong> TM 5 and a<br />

tryptophan W1.67 <strong>of</strong> TM 4 (Figure 9) (Wieland, Laak et al. 1999). Removing <strong>the</strong> hydrophobic<br />

phenylalanine at position 6.52 and 6.55 <strong>of</strong> <strong>the</strong> H1R showed a dramatic loss <strong>of</strong> affinity <strong>of</strong><br />

[ 3H]mepyramine, whereas replacing F6.55 and F6.52 with <strong>the</strong> hydrophobic but aliphatic<br />

23 | P a g e


Chapter 1<br />

methionine residue only slightly affected <strong>the</strong> binding <strong>of</strong> [ 3H]mepyramine, indicating that <strong>the</strong>se two<br />

phenylalanine residues in TM6 are most likely involved in a hydrophobic interaction with <strong>the</strong> H1R<br />

antagonist (Wieland, Laak et al. 1999).<br />

Active-state H1R<br />

It is believed that several highly conserved residues are involved in <strong>the</strong> activation mechanisms <strong>of</strong><br />

<strong>the</strong> rhodopsin family <strong>of</strong> GPCRs, including <strong>the</strong> H1R. Bakker and co-workers (Bakker, Jongejan et al.<br />

2008) have proposed that a conserved D2.50 <strong>of</strong> <strong>the</strong> NLxxxD motif in TM2 <strong>of</strong> <strong>the</strong> H1R forming a<br />

hydrogen bond network with W6.48 <strong>of</strong> <strong>the</strong> CWxP motif in TM6, in <strong>the</strong> same manner as in rhodopsin<br />

(Pardo, Deupi et al. 2007). This network constrains <strong>the</strong> toggle switch W6.48 in its inactive g+<br />

conformation (Ruprecht, Mielke et al. 2004) and maintains <strong>the</strong> H1R in <strong>the</strong> inactive conformation<br />

(Jongejan, Bruysters et al. 2005; Jongejan and Leurs 2005). N7.49 <strong>of</strong> <strong>the</strong> highly conserved NPxxY<br />

motif in TM7 <strong>of</strong> <strong>the</strong> H1R is restrained toward TM6 in <strong>the</strong> inactive g+ conformation (Govaerts, Lefort<br />

et al. 2001) by a water molecule (Bakker, Jongejan et al. 2008). The salt bridge interaction between<br />

R3.50 <strong>of</strong> <strong>the</strong> highly conserved E/DRY motif in TM3 with its adjacent D or E3.49 residue and an<br />

additional D/E 6.30 in TM6 (Alewijnse, Timmerman et al. 2000). Upon agonist binding, <strong>the</strong><br />

conformation <strong>of</strong> W6.48 was modified to point toward TM5. This occurs in <strong>the</strong> same concert manner<br />

with <strong>the</strong> side chain at position S3.36 in TM3 <strong>of</strong> <strong>the</strong> H1R (Jongejan, Bruysters et al. 2005; Jongejan<br />

and Leurs 2005). These conformation transitions <strong>of</strong> S3.36 and W6.48 disrupt <strong>the</strong> conserved<br />

hydrogen bond network between W6.48 and D2.50 and consequently trigger N7.49 to adopt <strong>the</strong><br />

active (trans) conformation to interact with D2.50 (Jongejan, Bruysters et al. 2005; Jongejan and<br />

Leurs 2005; Bakker, Jongejan et al. 2008). The conformational change <strong>of</strong> W6.48 also decrease <strong>the</strong><br />

proline kink <strong>of</strong> TM6, leading to <strong>the</strong> movement <strong>of</strong> <strong>the</strong> cytoplasmic end <strong>of</strong> TM6 away from TM3 and<br />

disrupting <strong>the</strong> ionic interaction between R3.50 with D3.49 and E6.30 (Shi, Liapakis et al. 2002).<br />

Bakker and co-workers also proposed that <strong>the</strong> physicochemical properties <strong>of</strong> <strong>the</strong> amino acid side<br />

chain at position 6.40 in TM6 are important in <strong>the</strong> process <strong>of</strong> receptor activation (Bakker, Jongejan<br />

et al. 2008). This position is located midway between <strong>the</strong> NPxxY motif and <strong>the</strong> ionic lock and it<br />

seems to be a highly conserved hydrophobic residue in <strong>the</strong> rhodopsin family (L, 14%; V, 42%, I,<br />

28% and M 5%). In <strong>the</strong> H1R, removal <strong>of</strong> <strong>the</strong> hydrophobic side chain (mutation I6.40A and I6.40G)<br />

and introduction <strong>of</strong> positive charge (mutation I6.40R and I6.40K) or negative charge (mutation<br />

I6.40E) resulted in <strong>the</strong> high constitutively active mutants (Bakker, Jongejan et al. 2008).<br />

24 | P a g e


<strong>Aim</strong> <strong>of</strong> <strong>the</strong> <strong>the</strong>sis<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


26 | P a g e<br />

<strong>Aim</strong> <strong>of</strong> <strong>the</strong> <strong>the</strong>sis<br />

Structure-based design is a powerful approach and a method <strong>of</strong> choice for lead optimization in drug<br />

discovery, if X-ray structures <strong>of</strong> a target protein are available. Structure-based design is widely used<br />

in drug discovery for soluble proteins, because <strong>the</strong>y are straightforward to isolate, purify and<br />

crystallize in comparisons to membrane proteins. Despite <strong>the</strong> importance <strong>of</strong> membrane proteins<br />

such as GPCRs as drug targets, <strong>the</strong>ir structure determination has been very limited compared to<br />

soluble proteins. The low success rate is due to <strong>the</strong> difficulties in obtaining large amounts <strong>of</strong> high<br />

quality protein. Additionally, <strong>the</strong> purification <strong>of</strong> GPCRs also requires detergents, which <strong>of</strong>ten have a<br />

negative effect on <strong>the</strong> protein stability and are also unfavorable during <strong>the</strong> crystallization process.<br />

The main aim <strong>of</strong> <strong>the</strong> research described in this PhD <strong>the</strong>sis is to learn more about <strong>the</strong> structural<br />

features <strong>of</strong> <strong>the</strong> H1R, a prototypical family A GPCR. In general, we employed approaches such as mass<br />

spectrometry (MS) and site-directed mutagenesis analysis based on molecular modeling <strong>of</strong> <strong>the</strong> H1R.<br />

The introduction to protein biochemistry, signal transduction pathways, and structural insight <strong>of</strong><br />

<strong>the</strong> family A <strong>of</strong> GPCRs and <strong>the</strong> H1R are discussed in Chapter 1. The expression and purification<br />

strategies <strong>of</strong> GPCRs and <strong>the</strong> H1R is fur<strong>the</strong>r discussed in Chapter 2. In Chapter 3, we show <strong>the</strong> use <strong>of</strong><br />

Baculovirus insect cell and <strong>the</strong> in vitro cell free expression systems for <strong>the</strong> production <strong>of</strong> <strong>the</strong> H1R<br />

protein, followed by <strong>the</strong> purification and <strong>the</strong> MS analysis <strong>of</strong> <strong>the</strong> H1R. We employed <strong>the</strong> method<br />

described for <strong>the</strong> protein production and MS analysis <strong>of</strong> <strong>the</strong> H1R subsequently for <strong>the</strong> human<br />

histamine H4 receptor (H4R). Using MS technology, we fur<strong>the</strong>r unraveled one <strong>of</strong> <strong>the</strong> glycosylation<br />

sites <strong>of</strong> <strong>the</strong> H1R in Chapter 4, and investigated if this proteomic approach can be useful in <strong>the</strong> study<br />

<strong>of</strong> ligand-receptor interaction. A preliminary result <strong>of</strong> covalent binding <strong>of</strong> an irreversible ligand<br />

(phenoxybenzamine) to <strong>the</strong> H1R is presented in this chapter. Using mutagenesis toge<strong>the</strong>r with<br />

homology modeling studies, we showed <strong>the</strong> importance <strong>of</strong> <strong>the</strong> highly conserved TAS (I/V) motif in<br />

TM3 and <strong>the</strong> residue I6.45 in <strong>the</strong> activation process <strong>of</strong> <strong>the</strong> H1R in Chapter 5 and Chapter 6,<br />

respectively. Finally, Chapter 7 summarizes <strong>the</strong> conclusions <strong>of</strong> <strong>the</strong> work presented in this <strong>the</strong>sis<br />

and <strong>the</strong> possible future experiments and developments are discussed.


CHAPTER<br />

GPCRs expression and purification<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Chapter 2<br />

28 | P a g e<br />

Introduction to <strong>the</strong> expression and purification <strong>of</strong> GPCRs<br />

Considering <strong>the</strong> crucial roles <strong>of</strong> G protein-coupled receptors (GPCRs) in many cellular and<br />

physiological processes, <strong>the</strong> knowledge regarding <strong>the</strong>ir 3D structural details will facilitate <strong>the</strong><br />

understanding <strong>of</strong> how GPCRs function at atomic level (Pierce, Premont et al. 2002; Horn, Bettler et<br />

al. 2003; Akermoun, Koglin et al. 2005; Lundstrom 2005). However, due to <strong>the</strong> low expression in<br />

native tissues and <strong>the</strong> difficulties in purification and crystallization process <strong>of</strong> GPCRs (Sarramegna,<br />

Talmont et al. 2003; Lundstrom, Wagner et al. 2006), <strong>the</strong> structural information <strong>of</strong> <strong>the</strong>se important<br />

membrane proteins has been limited for decades to crystal structures <strong>of</strong> <strong>the</strong> naturally abundant<br />

GPCR, rhodopsin (Palczewski, Kumasaka et al. 2000; Edwards, Li et al. 2004). Although, this<br />

rhodopsin crystals could provide significant advancements in GPCRs homology modeling, <strong>the</strong>y are,<br />

however, a light activated GPCR which somewhat differs from <strong>the</strong> o<strong>the</strong>r GPCRs that are activated by<br />

ligands (Topiol and Sabio 2008). Recently, new crystal structures <strong>of</strong> GPCRs; <strong>the</strong> β1AR (Warne,<br />

Serrano-Vega et al. 2008), <strong>the</strong> β2AR (Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et al.<br />

2007), <strong>the</strong> A2A adenosine receptor (Jaakola, Griffith et al. 2008) and <strong>the</strong> opsin (Park, Scheerer et al.<br />

2008; Scheerer, Park et al. 2008)) have finally been obtained. These structures will provide <strong>the</strong><br />

opportunity for evaluating an actual ligand-mediated GPCR in structure based design. However,<br />

much more high resolution information is needed, for example <strong>the</strong> structures <strong>of</strong> GPCRs from<br />

different families, structures bound with different ligands, and structures <strong>of</strong> GPCRs in complex with<br />

G proteins and o<strong>the</strong>r signaling molecules (Kobilka and Schertler 2008). In this chapter, we will<br />

discuss many aspects regarding <strong>the</strong> production <strong>of</strong> GPCRs in various heterologous expression<br />

systems, <strong>the</strong> techniques used in <strong>the</strong> purification and reconstitution processes towards<br />

crystallography.<br />

Recombinant expression <strong>of</strong> GPCRs<br />

With <strong>the</strong> exception <strong>of</strong> rhodopsin, GPCRs express naturally at low level (concentration range <strong>of</strong> 20-<br />

100 fmol/mg protein) (Lefkowitz 2007). Therefore, <strong>the</strong>y need to be heterologously over-expressed<br />

in <strong>the</strong> hosts such as Escherichia coli (E.coli), yeast, insect cells, or mammalian cells. The advantages<br />

and disadvantages <strong>of</strong> each expression system are summarized in Table 1. It is generally difficult to<br />

predict which expression system is best for <strong>the</strong> high level production <strong>of</strong> <strong>the</strong> targeted GPCRs. The<br />

right combination between <strong>the</strong> target receptor and expression host must be found most likely by<br />

trial and error (Grisshammer 2006; Wagner, Bader et al. 2006; McCusker, Bane et al. 2007).


GPCRs expression and purification<br />

However, as shown in Table 2, it is now possible to generate sufficient quantities <strong>of</strong> several GPCRs<br />

for crystallization or o<strong>the</strong>r biophysical studies purpose.<br />

Table 1<br />

Advantages and disadvantages <strong>of</strong> each expression systems<br />

Expression<br />

systems<br />

Advantages Disadvantages<br />

E.coli Rapid construction generation<br />

No native G-proteins<br />

Inexpensive<br />

Problems folding membrane proteins<br />

Ease <strong>of</strong> use<br />

Lack <strong>of</strong> post-translational modifications<br />

Simple culture requirements<br />

Non-native lipid environment<br />

Rapid doubling time<br />

Inclusion bodies require proper re-<br />

Variety <strong>of</strong> different strains (e.g. DH5α,<br />

BL21, CAG627, KS474)<br />

Produce inclusion bodies<br />

Possibility in using labeled amino acids or<br />

stable isotopes<br />

folding<br />

Yeast Rapid construction generation<br />

Non-native glycosylation<br />

Inexpensive<br />

Simple culture requirements<br />

Extensive genetics<br />

Eukaryotic processing<br />

Non-native lipid environment<br />

Insect cells More native lipid environment than yeast Cost<br />

Large scale culture applicable<br />

Non-native glycosylation<br />

Good safety pr<strong>of</strong>ile<br />

Non-native lipid environment<br />

Time required to generate virus<br />

Mammalian cells Native lipid environment<br />

Cost<br />

Native post-translational modifications Time required to establish stable cell<br />

Stable cell lines provide consistent lines<br />

expression level<br />

Transient transfected cells required large<br />

Transient transfected cells is fast<br />

amount <strong>of</strong> DNA<br />

Cell free Simple and Fast<br />

Cost<br />

High expression<br />

Open system<br />

Lack <strong>of</strong> post-translational modifications<br />

Prokaryotic: Escherichia coli (E.coli)<br />

E.coli is <strong>the</strong> most commonly used expression system. The major advantages <strong>of</strong> <strong>the</strong> E.coli expression<br />

system are its low cost, ease <strong>of</strong> use, variety <strong>of</strong> different strains, its rapid doubling time and a straight<br />

forward incorporation <strong>of</strong> specific labeled amino acids or stable isotopes. As a result <strong>of</strong> <strong>the</strong>se<br />

features, large scale protein production can be easily achieved (McCusker, Bane et al. 2007; Midgett<br />

29 | P a g e


Chapter 2<br />

and Madden 2007). A number <strong>of</strong> high-resolution structures have been obtained for soluble proteins<br />

recombinantly expressed in bacteria; however, it has been less successful in <strong>the</strong> case <strong>of</strong> membrane<br />

proteins such as GPCRs. Because <strong>of</strong> <strong>the</strong>ir transmembrane topology, membrane proteins are highly<br />

toxic to <strong>the</strong> bacterial host cells and <strong>the</strong>refore restricted <strong>the</strong>ir growth and reducing <strong>the</strong> protein yields<br />

(Lundstrom 2007). Heterologous expression <strong>of</strong> eukaryotic membrane proteins can also be<br />

hampered by a different syn<strong>the</strong>sis, targeting, insertion and folding characteristics in <strong>the</strong> prokaryotic<br />

host. For example, <strong>the</strong> polypeptide elongation and protein folding rates are considerably higher in<br />

prokaryotes than in eukaryotes, which can cause mistargeting and misfolding <strong>of</strong> heterologously<br />

expressed membrane proteins (Tate 2001). Also <strong>the</strong> lack <strong>of</strong> post-translation modifications in E.coli<br />

may provide both pros and cons. Homogenous protein is more suitable for crystallization, but some<br />

GPCRs (e.g. somatostatin, rhodopsin, β2AR, human thyrotropin and <strong>the</strong> bombesin BB2 receptors)<br />

requires glycosylation for high affinity ligand binding and/or G-protein coupling (Rands, Candelore<br />

et al. 1990; Rens-Domiano and Reisine 1991; Russo, Chazenbalk et al. 1991; Kaushal, Ridge et al.<br />

1994; Kusui, Benya et al. 1994) and o<strong>the</strong>rs (e.g. angiotensin II type 1 and follicle stimulating<br />

hormone receptors) require glycosylation for transport to <strong>the</strong> plasma membrane (Davis, Liu et al.<br />

1995; Deslauriers, Ponce et al. 1999; Jayadev, Smith et al. 1999; Dong, Filipeanu et al. 2007).<br />

However, it has been shown also that eukaryotic membrane proteins that are glycosylated in<br />

mammalian cells, such as <strong>the</strong> human CB2 receptor can still be functionally over-expressed in E.coli<br />

(Yeliseev, Wong et al. 2005). Ano<strong>the</strong>r disadvantage <strong>of</strong> <strong>the</strong> E.coli system is <strong>the</strong> difference in <strong>the</strong><br />

composition <strong>of</strong> <strong>the</strong> membranes between prokaryotes and eukaryotes. For example, <strong>the</strong> expression<br />

<strong>of</strong> <strong>the</strong> mammalian serotonin transporter and <strong>the</strong> binding properties <strong>of</strong> oxytocin and human μ-<br />

opioid receptors are dependent on cholesterol, which is not present in prokaryotic membranes<br />

(Hasegawa, Loh et al. 1987; Gimpl, Klein et al. 1995; Tate 2001). However, <strong>the</strong> addition <strong>of</strong> certain<br />

lipids during and/or after purification can sometimes restore functionality <strong>of</strong> over-expressed<br />

membrane proteins (Bogdanov, Umeda et al. 1999). Never<strong>the</strong>less, E.coli has been successfully used<br />

to express active GPCRs by implementing a combination <strong>of</strong> several strategies. The fusion <strong>of</strong> <strong>the</strong><br />

E.coli periplasmic maltose binding (PMB) protein to <strong>the</strong> N-terminus <strong>of</strong> GPCRs helps targeting <strong>the</strong><br />

expressed protein to <strong>the</strong> periplasmic membrane (which corresponds to <strong>the</strong> outer side <strong>of</strong> a<br />

eukaryotic plasma membrane) as reported for <strong>the</strong> human adenosine A2A, rat neurotensin, human<br />

serotonin 5HT1a, rat NK2, muscarinic M1 and M2, human β2AR and <strong>the</strong> human cannabinoid CB2<br />

receptors (Bertin, Freissmuth et al. 1992; Grisshammer, Little et al. 1994; Hulme and Curtis 1998;<br />

Furukawa and Haga 2000; Hampe, Voss et al. 2000; Weiss and Grisshammer 2002; White, Trinh et<br />

al. 2004; Yeliseev, Wong et al. 2005). Also, <strong>the</strong> use <strong>of</strong> o<strong>the</strong>r fusion proteins such as a Bacillus subtilis<br />

30 | P a g e


GPCRs expression and purification<br />

integral membrane protein (Mistic) (Roosild, Greenwald et al. 2005), and thioredoxin (TrxA)<br />

(Tucker and Grisshammer 1996) have been reported to be a useful fusion protein for GPCR<br />

expression. Ano<strong>the</strong>r approach has been to produce large amounts <strong>of</strong> eukaryotic membrane proteins<br />

as inclusion bodies in E.coli. The protein in inclusion bodies is more resistant to proteolytic<br />

degradation and easier to purify. However, to successfully obtain GPCRs with this approach, a<br />

proper re-folding <strong>of</strong> protein is required. This technique has only been successful for <strong>the</strong> olfactory<br />

OR5 receptor, <strong>the</strong> human leukotriene B4 receptor and <strong>the</strong> human serotonin 5HT4a receptor (Kiefer,<br />

Krieger et al. 1996; Baneres, Martin et al. 2003; Baneres, Mesnier et al. 2005).<br />

O<strong>the</strong>r prokaryotic hosts<br />

In addition to E.coli, <strong>the</strong>re are several o<strong>the</strong>r bacterial expression systems for <strong>the</strong> production <strong>of</strong><br />

GPCRs. Unlike in E.coli system, <strong>the</strong> protein expressed in a gram-positive bacteria Lactococcus lactis<br />

could be transported to <strong>the</strong> cytoplasmic membrane, <strong>the</strong>refore, circumventing <strong>the</strong> forming <strong>of</strong><br />

inclusion bodies (Monne, Chan et al. 2005). The L. lactis has been used successfully for <strong>the</strong><br />

expression <strong>of</strong> prokaryotic proteins such as <strong>the</strong> ABC transporter, and <strong>the</strong> yeast mitochondrial carrier<br />

proteins CTP1 and AAC3 (Kunji, Slotboom et al. 2003). However, <strong>the</strong> expression <strong>of</strong> eukaryotic<br />

membrane protein such as <strong>the</strong> human KDEL receptor in L. lactis only resulted in a yield less than<br />

0.1% <strong>of</strong> <strong>the</strong> total membrane protein (de Ruyter, Kuipers et al. 1996). Bacillus subtilis, ano<strong>the</strong>r Gram-<br />

positive bacterium, has been considered as a good host for expressing large quantities <strong>of</strong><br />

recombinant interferon (Palva, Lehtovaara et al. 1983) and human cystatins (Akiba, Hayashi et al.<br />

2006). However, its high endogenous protease activity has limited <strong>the</strong> use <strong>of</strong> this system<br />

(Lundstrom 2007). Halinobacterium salinarum has become an alternative prokaryotic expression<br />

host. Various membrane proteins such as <strong>the</strong> E.coli aspartate transcarbamylase, <strong>the</strong> yeast α mating<br />

factor receptor, <strong>the</strong> human muscarinic M1, <strong>the</strong> serotonin 5-HT2 receptor and <strong>the</strong> adrenergic α2B<br />

receptor have been expressed from H. salinarum (Turner, Reusch et al. 1999; Bartus, Jaakola et al.<br />

2003). However, <strong>the</strong> GPCRs were expressed at low level. Interestingly, <strong>the</strong> expressed adrenergic α2B<br />

receptor showed functional binding activity, although with 10 times lower binding affinities than in<br />

mammalian cells (Bartus, Jaakola et al. 2003). Recently, due to its advantage in providing more<br />

membrane surface per cell compared to o<strong>the</strong>r typical expression host, <strong>the</strong> photosyn<strong>the</strong>tic<br />

bacterium Rhodobacter sphaeroides has also been introduced to express some GPCRs such as <strong>the</strong><br />

human adenosine A2a, <strong>the</strong> human angiotensin AT1a and <strong>the</strong> human bradykinin B2 receptors (Roy,<br />

Shukla et al. 2008).<br />

31 | P a g e


Chapter 2<br />

Yeast<br />

Yeasts strains are valuable candidates for heterologous protein expression. Like E.coli, <strong>the</strong>y are well<br />

characterized, easy to culture at low cost and straightforward to manipulate genetically. An<br />

important feature <strong>of</strong> yeast cells is also that <strong>the</strong>y possess a G protein which activates a MAP kinase<br />

cascade, similar to that found in mammalian cells. Therefore, heterologous receptors expressed in<br />

yeast cells can be coupled to this endogenous signaling pathway, potentially leading to a functional<br />

response (Francken, Luyten et al. 1997; Midgett and Madden 2007). Fur<strong>the</strong>rmore, yeasts are<br />

eukaryotic cells and have <strong>the</strong> potential to perform most types <strong>of</strong> post-translational modifications<br />

and form disulfide bonds, similar to more complex eukaryotic cells. Yet, glycosylation is mostly <strong>of</strong><br />

<strong>the</strong> high-mannose type and appears limited for heterologous membrane proteins (Eckart and<br />

Bussineau 1996; Midgett and Madden 2007). However, yeast with humanized glycosylation<br />

machineries has also been introduced and may provide an important alternative expression system<br />

(Hamilton, Davidson et al. 2006). It is also worthy to note that yeast cells are surrounded by a cell<br />

wall that may hinder <strong>the</strong> recovery <strong>of</strong> non-secreted proteins such as GPCRs (Francken, Luyten et al.<br />

1997; Midgett and Madden 2007). Moreover, <strong>the</strong> low cholesterol content <strong>of</strong> <strong>the</strong> yeast membranes<br />

may result in altered ligand binding properties <strong>of</strong> some GPCRs, although, co-solubilization with<br />

mammalian lipids such as cholesterol has been shown to stabilize <strong>the</strong> active conformation <strong>of</strong> <strong>the</strong> μ-<br />

opioid receptor (Opekarova and Tanner 2003). The yeast systems that are widely used for<br />

heterologous protein expressions are Saccharomyces cerevisiae (S. cerevisiae), Schizosaccharomyces<br />

pombe (S. pombe) and Pichia pastoris (P. pastoris). The S. cerevisiae or baker’s yeast has been used<br />

extensively to express GPCRs, for example, <strong>the</strong> human A2A receptor fused with C-terminal green<br />

fluorescent has been highly expressed (~4 mg functional protein per litre <strong>of</strong> culture) in S. cerevisiae<br />

(Niebauer and Robinson 2006). The S. pombe or fission yeast is less commonly used. Yet, it has<br />

recently been examined as it showed <strong>the</strong> ability to express higher complexity eukaryotic proteins<br />

(Giga-Hama and Kumagai 1998). Comparison <strong>of</strong> GPCR expression demonstrated that <strong>the</strong> human<br />

dopamine D2 receptor was expressed at 5-fold higher levels in <strong>the</strong> S. pombe than in <strong>the</strong> S. cerevisiae<br />

(Sander, Grunewald et al. 1994). The methylotrophic yeast P. pastoris is capable <strong>of</strong> achieving higher<br />

cell densities than <strong>the</strong> yeast S. cerevisiae, <strong>the</strong>refore, allowing production <strong>of</strong> g/L quantities <strong>of</strong><br />

recombinant proteins. 100 GPCRs were expressed in P. pastoris, for which a large number <strong>of</strong> GPCRs<br />

could reach a high expression level (1-10 mg/L). The highest level (180 pmol/mg receptor protein)<br />

was measured for <strong>the</strong> human adenosine A2a receptor (Lundstrom, Wagner et al. 2006). It has also<br />

been shown that <strong>the</strong> expression <strong>of</strong> <strong>the</strong> human opioid receptor in P. pastoris could be improved by<br />

32 | P a g e


GPCRs expression and purification<br />

fusing <strong>the</strong> S. cerevisiae yeast α-factor signal sequence to <strong>the</strong>ir N-terminal tails (Sarramegna, Muller<br />

et al. 2005).<br />

Insect cells<br />

The use <strong>of</strong> baculovirus vectors for <strong>the</strong> infection <strong>of</strong> insect cell lines is <strong>the</strong> second most commonly<br />

used expression system aster E.coli. The baculovirus expression system (BVES) emerged from<br />

studies designed to control insect pests (Smith, Summers et al. 1983). However, <strong>the</strong> system has<br />

gained more widespread use as a method for protein expression. The BVES requires <strong>the</strong><br />

introduction <strong>of</strong> <strong>the</strong> foreign GPCR gene <strong>of</strong> interest into a non-essential region <strong>of</strong> <strong>the</strong> viral genome,<br />

usually Autographa california nuclear polyhedrosis virus (AcNPV), by homologous recombination<br />

with a transfer plasmid (McCusker, Bane et al. 2007). The production <strong>of</strong> protein is produced when<br />

insect cells are infected with <strong>the</strong> recombinant virus (Fraser 1992). Most <strong>of</strong> <strong>the</strong> cells that are<br />

commonly used are derived from ovarian tissue <strong>of</strong> <strong>the</strong> fall army worm Spodoptera frugiperda (Sf9<br />

and Sf21). Although <strong>the</strong> E.coli and yeast systems are less expensive to culture, <strong>the</strong> advent <strong>of</strong> serum-<br />

free media for growing <strong>of</strong> Sf9 or Sf21 cell lines and <strong>the</strong> use <strong>of</strong> intermediate-scale shaker suspension<br />

have made <strong>the</strong> insect cells more widely accessible (Midgett and Madden 2007). Insect cells provide<br />

a correct folding <strong>of</strong> recombinant protein as well as disulfide bond formation and perform post<br />

translational modifications (such as fatty acid acylation and phosphorylation) identical to those in<br />

mammalian cells, except for glycosylation. The oligosaccharides produced in insect cells are smaller<br />

and <strong>of</strong> <strong>the</strong> high-mannose type. However, a characterization <strong>of</strong> a new insect cell line, Trichoplusia ni<br />

(Tni, High Five), showed that this cell line is able to produce sugars <strong>of</strong> <strong>the</strong> complex type (Ogonah,<br />

Freedman et al. 1995). Sf9 cells could also express certain Gα subunits indicating <strong>the</strong>ir ability to<br />

provide favorable interactions for GPCRs (Kleymann, Boege et al. 1993; Mulheron, Casanas et al.<br />

1994; Obosi, Schuette et al. 1996; Kuhn and Gudermann 1999). Consequently, GPCRs that are<br />

expressed in insect cells system are almost always able to bind to <strong>the</strong>ir respective ligands<br />

(McCusker, Bane et al. 2007). A drawback to <strong>the</strong> baculovirus system is <strong>the</strong> requirement for high<br />

titer viral stocks to infect cell cultures. Productions <strong>of</strong> <strong>the</strong>se stocks are time-consuming and<br />

laborious to perform (McCusker, Bane et al. 2007). Ano<strong>the</strong>r issue is <strong>the</strong> quality <strong>of</strong> protein crystal<br />

because insect cells <strong>of</strong>ten give heterogeneous populations <strong>of</strong> receptors due to variable post-<br />

translational modifications (Reilander, Boege et al. 1991). Moreover, insect plasma membranes are<br />

not identical to those <strong>of</strong> mammalian cells. They have higher amounts <strong>of</strong> unsaturated lipids and<br />

lower amount <strong>of</strong> cholesterol (Marheineke, Grunewald et al. 1998), which can alter ligand binding<br />

(Gimpl, Klein et al. 1995). However, once <strong>the</strong> tedious start up process has finished, <strong>the</strong> insect cell<br />

33 | P a g e


Chapter 2<br />

system is easy to scale up and has produced numerous GPCRs (as shown in Table 1) (Grisshammer<br />

and Tate 1995; Massotte 2003; Ratnala, Swarts et al. 2004; Akermoun, Koglin et al. 2005; Cherezov,<br />

Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007; Sansuk, Balog et al. 2008; Warne, Serrano-<br />

Vega et al. 2008). After several attempts, <strong>the</strong> crystals <strong>of</strong> ligand-mediated GPCR have finally been<br />

obtained by using insect cells as an expression system. The β2AR crystals were made by forming a<br />

complex with an antibody Fab fragment (Rasmussen, Choi et al. 2007) or by fusing in <strong>the</strong> middle <strong>of</strong><br />

<strong>the</strong> third ICL3 to T4 lysozyme (Cherezov, Rosenbaum et al. 2007). The turkey β1AR was mutated to<br />

enhance <strong>the</strong>rmo stability over <strong>the</strong> wild type receptor and finally yielded a good crystal <strong>of</strong> 2.7Å<br />

(Warne, Serrano-Vega et al. 2008) and <strong>the</strong> A2A adenosine receptor protein was stabilized through<br />

<strong>the</strong> T4 lysozyme insertion (Jaakola, Griffith et al. 2008).<br />

Mammalian cells<br />

Among heterologous expression systems, mammalian expression systems, although costly and<br />

time-consuming, clearly <strong>of</strong>fers <strong>the</strong> most native cellular environment for expression <strong>of</strong> GPCRs.<br />

Mammalian cells contain necessary co-receptors, providing <strong>the</strong> optimal membrane composition<br />

and can perform <strong>the</strong> right post translational modifications (McCusker, Bane et al. 2007).<br />

Heterologous protein expression in mammalian cells may be achieved through transient or stable<br />

expression and <strong>the</strong> use <strong>of</strong> viral vectors. Recently, Standfuss and co-workers have reported a crystal<br />

structure <strong>of</strong> a rhodopsin mutant produced from transfected COS-7 cells. The final yield <strong>of</strong> purified<br />

protein is about 2.5 mg <strong>of</strong> rhodopsin from 50 plates (15 cm) <strong>of</strong> COS-7 cells (Standfuss, Xie et al.<br />

2007). Some mammalian cell lines such as HEK293 and CHO cells are able to be stably transfected<br />

and to grow in suspension. However, <strong>the</strong> expression levels obtained are <strong>of</strong>ten low and it is difficult<br />

to obtain reliable stable cell lines due to cell toxicity in <strong>the</strong> presence <strong>of</strong> high levels <strong>of</strong> GPCRs; for<br />

example <strong>the</strong> β2-adrenergic and μ-opioid receptors could not be maintained in suspension culture<br />

(Stanasila, Pattus et al. 1998; Kohno, Fukushima et al. 2000). To overcome <strong>the</strong>se limitations,<br />

inducible mammalian cell systems have been developed (Tate 2001; Neddermann, Gargioli et al.<br />

2003; Lipscomb, Mowry et al. 2004; Weber, Rimann et al. 2004; Ivanova, Brandli et al. 2005; Bach,<br />

Grigat et al. 2007). Recently, it was found that <strong>the</strong> tetracycline-inducible system enabled a codon-<br />

optimized hamster β2-adrenergic receptor to be expressed at levels <strong>of</strong> 220 ± 40 pmol/mg protein<br />

(Chelikani, Reeves et al. 2006). Ano<strong>the</strong>r approach to produce high levels <strong>of</strong> GPCRs is <strong>the</strong> use <strong>of</strong> viral<br />

vectors, such as <strong>the</strong> Semliki Forest Virus (SFV), to infect mammalian host cells. In many cases,<br />

relatively high yields <strong>of</strong> protein have been obtained, such as adenosine A2A receptors with <strong>the</strong><br />

expression level <strong>of</strong> 287 pmol/mg protein (Hassaine, Wagner et al. 2006). Besides SFV, it has also<br />

34 | P a g e


GPCRs expression and purification<br />

been shown that baculoviruses are able to infect mammalian cells (H<strong>of</strong>mann, Sandig et al. 1995;<br />

Shoji, Aizaki et al. 1997). This approach has recently been used for large scale protein production in<br />

mammalian cell expression system (Scott, Modha et al. 2007) by employing <strong>the</strong> pfastbacmam-1<br />

vector to generate a recombinant baculovirus in which <strong>the</strong> polyhedrin promoter is replaced with a<br />

cytomegalovirus (CMV) promoter (Condreay, Wi<strong>the</strong>rspoon et al. 1999). Both <strong>the</strong> generation <strong>of</strong><br />

stable cells containing GPCRs <strong>of</strong> interest and <strong>the</strong> handling <strong>of</strong> viral stocks can be laborious and time<br />

consuming. As a result, with <strong>the</strong> development <strong>of</strong> serum-free media and <strong>of</strong> relatively cheap<br />

transfection reagents, large-scale transient Transfection (LSTT) <strong>of</strong> mammalian cells has been used<br />

as an alternative expression method (Derouazi, Girard et al. 2004; Geisse, Jordan et al. 2005; Baldi,<br />

Hacker et al. 2007).<br />

Cell Free Expression System<br />

Cell Free Expression is an alternative method that allows full control and high flexibility <strong>of</strong><br />

conditions to produce protein <strong>of</strong> interest without intact cells. The protein syn<strong>the</strong>sis can be<br />

performed ei<strong>the</strong>r in a batch reaction using a test tube or ano<strong>the</strong>r container with a fixed volume or<br />

in a continuous exchange reaction. In <strong>the</strong> batch mode, <strong>the</strong> translation <strong>of</strong> protein stops as soon as <strong>the</strong><br />

substrate is finished, or any product or by-product reaches an inhibitory concentration, usually<br />

within 20 to 60 minutes <strong>of</strong> incubation. Due to <strong>the</strong>se limitations (e.g. a short lifetime and low<br />

productivities for a preparative scale), a continuous cell free expression systems was first<br />

introduced in 1988 by Spirin and co-workers (Spirin, Baranov et al. 1988). The principle <strong>of</strong> this<br />

system was <strong>the</strong> continuous addition <strong>of</strong> reaction substrates (amino acids and nucleoside<br />

triphosphates) with continuous removal <strong>of</strong> reaction products through a porous membrane during<br />

incubation. This design could produces hundreds <strong>of</strong> micrograms per ml reaction volume. Recently<br />

several modified versions <strong>of</strong> <strong>the</strong> Spirin system have been reported (Kigawa and Yokoyama 1991;<br />

Kawarasaki, Kawai et al. 1995; Kim, Kigawa et al. 1996; Kigawa, Yabuki et al. 1999). Kigawa and co-<br />

workers showed that by using a dialysis membrane to facilitate <strong>the</strong> continuous supply <strong>of</strong> substrate<br />

and removal <strong>of</strong> by-products, an E.coli coupled transcription-translation system yields as much as 6<br />

mg protein per ml reaction volume (Kigawa, Yabuki et al. 1999). Besides <strong>the</strong> feasibility to produce<br />

milligram <strong>of</strong> proteins, Cell Free Expression system also <strong>of</strong>fers a solution for traditional in vivo<br />

expression problems such as cell-induced protein degradation and aggregation, loss <strong>of</strong> DNA<br />

template, and toxicity (Kigawa, Yabuki et al. 1999). However, this system can be costly and it lacks<br />

GPCR-interacting signaling components and post translational modifications, resulting in <strong>the</strong><br />

misfolding <strong>of</strong> proteins (McCusker, Bane et al. 2007). The E.coli cell free expression system was<br />

35 | P a g e


Chapter 2<br />

successfully used for stable-isotope labeling <strong>of</strong> proteins for NMR spectroscopy such as amino acid-<br />

selective 15N labeling (Kigawa, Muto et al. 1995), dual-amino acid selective 13C- 15N labeling (Yabuki,<br />

Kigawa et al. 1998), and uniform 13C/ 15N-labeling (Kigawa, Yabuki et al. 1999). Ano<strong>the</strong>r cell free<br />

expression system has also been developed from <strong>the</strong> wheat embryos. They possess numerous<br />

advantages such as low cost, easy availability in large amounts, and low endogenous incorporation<br />

(Madin, Sawasaki et al. 2000). This system has been applied for high throughput functional<br />

proteomics (Sawasaki, Ogasawara et al. 2002). More than 50 genes can be translated in parallel,<br />

yielding between 0.1-2.3 mg <strong>of</strong> protein in 2 days. The wheat germ cell free expression system also<br />

<strong>of</strong>fers <strong>the</strong> capacity to syn<strong>the</strong>size high molecular weight proteins and is <strong>the</strong>refore more suitable for<br />

<strong>the</strong> expression <strong>of</strong> eukaryotic proteins. However, up to now expression <strong>of</strong> GPCRs using cell free<br />

systems is still limited and depends on fusion partners. By adding an N-terminal thioredoxin tag,<br />

<strong>the</strong> human β2 adrenergic, rat muscarinic acetylcholine, rat neurotensin and human histamine H1R<br />

have been expressed using E.coli cell free expression systems (Ishihara, Goto et al. 2005; Sansuk,<br />

Balog et al. 2008). A N-terminal T7 tag was essential for expression <strong>of</strong> <strong>the</strong> porcine vasopressin<br />

receptor (Klammt, Schwarz et al. 2005). In all cases, GPCRs were produced as ei<strong>the</strong>r precipitates or<br />

detergent soluble protein by <strong>the</strong> addition <strong>of</strong> detergents to <strong>the</strong> reaction mixture. Expression levels<br />

have been reported as high as milligrams per milliliter <strong>of</strong> reaction mixtures. However, <strong>the</strong><br />

radioligand binding results <strong>of</strong> GPCRs produced from cell free expression system have been reported<br />

for only β2AR (Ishihara, Goto et al. 2005) and histamine H1R (Sansuk, Balog et al. 2008).<br />

Transgenic animals for protein production<br />

The first crystal structure <strong>of</strong> GPCR was successfully obtained by simply isolating rhodopsin from<br />

bovine retina(Palczewski, Kumasaka et al. 2000). This success took advantage <strong>of</strong> <strong>the</strong> naturally<br />

enriched levels <strong>of</strong> rhodopsin that occur in retina and <strong>the</strong> fact that this protein is uniformly<br />

glycosylated to provide large amounts <strong>of</strong> homogenous protein for crystallization. To fur<strong>the</strong>r apply<br />

<strong>the</strong>se advantages for <strong>the</strong> expression <strong>of</strong> o<strong>the</strong>r GPCRs, some laboratories have been exploited <strong>the</strong><br />

retina’s rod cells from several animal species as a “bioreactor” to produce nascent heterologous<br />

GPCRs for purification and crystallization. Pro<strong>of</strong> <strong>of</strong> this concept has been reported with <strong>the</strong> fruit fly<br />

(Drosophila melanogaster) (Eroglu, Cronet et al. 2002) , <strong>the</strong> African clawed toad ( Xenopus laevis)<br />

(Zhang, Salom et al. 2005), and <strong>the</strong> mouse (Li, Salom et al. 2007; Salom, Wu et al. 2008). The fruit fly<br />

D. melanogaster has a high biogenesis capacity for membrane proteins and sufficient space in <strong>the</strong><br />

membrane to accommodate <strong>the</strong> over-expressed protein (Wagner, Bader et al. 2006). They have<br />

been successfully used to over-express <strong>the</strong> D. melanogaster metabotropic glutamate receptor<br />

36 | P a g e


GPCRs expression and purification<br />

(Eroglu, Cronet et al. 2002). The use <strong>of</strong> X. laevis for protein expression is still limited to <strong>the</strong> small<br />

scale since it requires injection <strong>of</strong> in vitro syn<strong>the</strong>sized mRNA into each individual oocyte (Zhang,<br />

Salom et al. 2005). The attempts to replace endogenous retinal rhodopsin in mouse eyes with a<br />

heterologous GPCR such as adenosine A1 receptor and 5HT4 were reported (Li, Salom et al. 2007;<br />

Salom, Wu et al. 2008). However, <strong>the</strong> amount <strong>of</strong> receptor protein presents in <strong>the</strong> engineered retina<br />

was still less than what <strong>the</strong>y have expected. The additional efforts in improving homozygote lines<br />

and extraction-purification protocol are required (Li, Salom et al. 2007; Salom, Wu et al. 2008).<br />

Purification <strong>of</strong> GPCRs<br />

In general, <strong>the</strong> purification <strong>of</strong> membrane protein such as GPCRs is not as easy as <strong>the</strong> purification <strong>of</strong><br />

water-soluble proteins, although <strong>the</strong> methods used are generally <strong>the</strong> same. The first purification<br />

step is <strong>the</strong> preparation <strong>of</strong> a receptor-enriched fraction, followed by <strong>the</strong> solubilization <strong>of</strong> <strong>the</strong> GPCRs<br />

from <strong>the</strong> membrane compartment. Once suitable conditions for solubilization have been identified,<br />

a robust and effective purification scheme is required. Since purification <strong>of</strong> membrane proteins<br />

must be achieved in <strong>the</strong> presence <strong>of</strong> detergents in order to maintain <strong>the</strong> solubility <strong>of</strong> <strong>the</strong><br />

hydrophobic proteins, <strong>the</strong> application <strong>of</strong> standard chromatographic procedures is not so straight<br />

forward (Sarramegn, Muller et al. 2006). Charged steroid-based detergents, such as sodium cholate<br />

or deoxycholate, will prevent <strong>the</strong> binding <strong>of</strong> proteins to ion-exchange or dye matrix columns. Non-<br />

ionic detergents such as dodecyl maltoside which are more suitable in many cases are not<br />

compatible with <strong>the</strong> use <strong>of</strong> ammonium sulphate. However, once <strong>the</strong> protein is bound to a column, it<br />

is possible to exchange detergent simply by washing <strong>the</strong> column with a buffer containing a different<br />

detergent (Gwyn 1994). In addition to this, <strong>the</strong> number <strong>of</strong> column steps should be kept to a<br />

minimum to preserve receptor integrity at maximum yield (Grisshammer, White et al. 2005).<br />

Finally, in order to study membrane proteins in a stable environment, <strong>the</strong> solubilized and purified<br />

proteins need to be reconstituted back into an artificial environment that mimics <strong>the</strong>ir native<br />

membranes.<br />

The solubilization step<br />

The objective <strong>of</strong> <strong>the</strong> solubilization step is to extract membrane proteins from <strong>the</strong>ir lipid<br />

environment using detergents. Detergents are amphiphillic molecules that can disrupt <strong>the</strong> lipid<br />

bilayer by dispersing <strong>the</strong> lipids and membrane protein into detergent micelles (Sarramegna, Muller<br />

et al. 2005). As shown in Table 3, <strong>the</strong>re are various kinds <strong>of</strong> detergents. Non-ionic detergents<br />

37 | P a g e


Chapter 2<br />

contain uncharged hydrophilic head groups <strong>of</strong> ei<strong>the</strong>r polyoxyethylene or glycosidic groups (e.g.<br />

Triton X-100, DM, digitonin).<br />

Table 2<br />

Summary <strong>of</strong> <strong>the</strong> GPCRs protein production from various expression systems<br />

Expression<br />

System<br />

38 | P a g e<br />

Year References GPCRs Species Expression<br />

level<br />

Purified<br />

Protein<br />

E. coli 1988 Marullo, 1998 β2 Human 0.2% ND<br />

1992 Bertin, 1992 5-HT1A human ND ND<br />

1993 Haendler, 1993 ETB human ND ND<br />

1993 Grisshammer, 1993 NTS1 rat 15 pmol/mg ND<br />

1994 Ho, 2008 NK2 rat ND ND<br />

1994 Jockers, 1994 A1A human 0.4 pmol/mg ND<br />

1996 Tucker, 1996 NTS1 rat 15 pmol/mg ND<br />

1996 Kiefer, 1996 OR5 ND 10% 20 mg/L<br />

1997 Grisshammer, 1997 NTS1 rat 13 pmol/mg ND<br />

2000 Furukawa, 2000 M2 human 0.2nmol/0.1L ND<br />

2000 Hampe, 2000 β2 human 6 pmol/mg ND<br />

2001 Busuttil, 2001 TSH human ND 2.5 mg/L<br />

2002 Weiss, 2002 A2A human 20 nmol/L ND<br />

2003 Baneres, 2003 BLT1 human 10 mg/L 2-3 mg/L<br />

2004 White, 2004 NTS1 rat 0.45 mg/L 0.2 mg/L<br />

2005 Yeliseev, 2005 CB2 human 1-2 mg/L 1 mg/L<br />

2005 Baneres, 2005 5HT4 mouse 5-8 mg/L 0.5 mg/L<br />

2006 Krepkiy, 2006 CB2 human 15 pmol/mg ND<br />

2007 Harding, 2007 NTS1 rat ND ND<br />

S. cerevisiae<br />

S. pombe<br />

2007 Bane, 2007 NK1 human ND 5 mg/L<br />

1996 Mollaaghababa, 1996 Rho bovine 2 mg/L 40 μg/L<br />

1998 Andersen, 1998 D1 human 13 pmol/mg ND<br />

2001 Liitti, 2001<br />

α 2C human 25 pmol/mg ND<br />

2007 O’Malley, 2007 α2A Human 4 mg/L 1 mg/L<br />

1995 Ficca, 1995 Β2 human 7.5 pmol/mg ND<br />

1995 Arkinstall, 1995 NK2 human 1.2 pmol/mg ND<br />

1996 Talmont, 1996 Μ human 0.4 pmol/mg ND<br />

P. pastoris 2001 Schiller, 2001 ETB human 45 pmol/mg 15pmol/mg<br />

2002 Feng, 2002 CB2 human ND ND<br />

2004 De Jong, 2004 D2 human 13 pmol/mg ND<br />

2005 Kim, 2005 CB1 human 3.6 pmol/mg ND<br />

2005 Sarramegna, 2005 Μ human 30 mg/L 12 mg/L<br />

2006 Andre, 2006 DRD2 human 39.1 ND<br />

2006 Andre, 2006 NK1 human 24 ND


GPCRs expression and purification<br />

Table 2 (continued)<br />

Expression Year References GPCRs Species Expression<br />

System<br />

level<br />

2006 Andre, 2006<br />

NK2 human 120<br />

2006 Andre, 2006 5-HT5A human 39.1<br />

2006 Andre, 2006<br />

human 76<br />

Insect cells<br />

HEK293<br />

Α2A<br />

Purified<br />

Protein<br />

ND<br />

ND<br />

ND<br />

2007 Li, 2007 CB2 human ND ND<br />

2008 MAgnin, 2009 H1 human ND 0.5 mg/L<br />

1991 Parker, 1991 M1 human 2 nmol/l 0.3 nmol/L<br />

1991 Parker, 1991 β2 turkey 2 nmol/l 0.15 nmol.L<br />

1991 Parker, 1991 M2 Human 10-15 nmol/l 4.5-7 nmol/l<br />

1991 Reilander, 1991 β2 human 12-19 ND<br />

1993 Kwatra, 1993 β2 human 7 ND<br />

1993 Kwatra, 1993 NK1 rat 1-2 ND<br />

1995 Kwatra, 1995 β2 human 5-20 1-4 nmol/l<br />

1995 Kobilka, 1995 β2 human 2.5 5 nmol/l<br />

1996 Hayashi, 1996 M2 human 10.8 nmol/l 4.4 nmol/l<br />

1996 Nekrasova, 1996 OR4 human ND ND<br />

1998 Ohtaki, 1998 PAC1 human 50-150 250 µg/l<br />

1999 Klaassen, 1999 Rho bovine 4 mg/l 0.5-1 mg/l<br />

2000 Ghanouni, 2000 β2 human ND ND<br />

2000 Hampe, 2000 β2 human 17 ND<br />

2003 Warne, 2003 β1 turkey 7.5 mg/l 2.5 mg/L<br />

2003 Park, 2003 M2 human ND ND<br />

2003 Panneels, 2003 Glu Dros 11 ND<br />

2003 Bodor, 2003 P2Y12 human ND 0.3-0.5 mg/L<br />

2004 Ratnala, 2004 H1R human 4-7 mg/l 3-4 mg/L<br />

2004 Asmann, 2004 Secretin human 830 1.4 mg/L<br />

2005 Akermoun, 2005 H1 human 85 pmol/mg ND<br />

2007 Rasmussen, 2007 β2 human ND 8-12 mg/L<br />

2007 Cherezov, 2007 β2 human ND 5 mg/L<br />

2008 Serrano-Vega, 2008 β1 turkey ND 2.5 mg/L<br />

2008 Jaakola, 2008 A2A human ND 3 mg/L<br />

2008 Ruan, 2008 TP human ND 4 mg/0.3 L<br />

2002 Reeves, 2002 Rho bovine 10 mg/l 9 mg/l<br />

2002 Vicentic, 2002 α1A, B, D human 0.15-3.6 ND<br />

2003 Christ<strong>of</strong>fers, 2003 μ mouse 10 ND<br />

2005 Christ<strong>of</strong>fers, 2005 Δ mouse 10 ND<br />

2006 Gan, 2006 PTH1 human ND ND<br />

2006 Chelikani, 2006 β2 human 50 μg/plate 12 μg/plate<br />

2007 Daulat, 2007 MT1-2 human ND ND<br />

2007 Standfuss, 2007 Rho human ND 3mg/50plate<br />

COS-7 2003 Sen, 2003 Α2B human 176 pmol/mg ND<br />

2007 Hassaine, 2006 Α2A human 287 pmol/mg ND<br />

39 | P a g e


Chapter 2<br />

Table 2 (continued)<br />

Expression Year References GPCRs Species Expression Purified<br />

System<br />

level Protein<br />

Cell free 2005 Ishihara, 2005 β2 human 1 mg/ml ND<br />

2007 Klammt, 2007 ETB human 3 mg/ml ND<br />

2007 Klammt, 2007 V2R human 3 mg/ml ND<br />

2008 Sansuk, 2008 H1 human 3-5 mg/ml 0.2 mg<br />

Drosophilla 2002 Eroglu, 2002 DmGluRA Dros 170µg/g fly ND<br />

melanogaster<br />

heads<br />

Xenopus 2005 Zhang, 2005 5-HT1A human 1-5<br />

ND<br />

laevis<br />

ng/tadpole<br />

Mouse 2007 Li, 2007 A1 Human 90 ng/eye ND<br />

2008 Salom, 2008 5-HT4 human 100 ng /eye ND<br />

Abbreviations: α,β, adrenergic receptor; 5-HT, hydroxytryptamine receptor; ET, endo<strong>the</strong>lin receptor; NTS,<br />

neurotensin receptor; NK, neurokinin receptor; OR, olfactive receptor; TSH, thyroid-stimulating hormone<br />

receptor; BLT, leukotriene receptor; CB, cannabinoid receptor; A, adenosine receptor; TP, tromboxane<br />

receptor; M, muscarinic receptor; D, dopamine receptor; Rho, rhodopsin; μ-, δ-opioid, opioid receptor; PAC,<br />

pituitary adenyl cyclase-activating polypeptide receptor; Glu, metabolic glutamate receptor; P2Y, purinergic<br />

receptor; MT, melatonin receptor; V, vasopressin receptor; PTHR, parathyroid hormone receptor; H,<br />

histamine receptor; DmGluRA, Drosophila metabotropic glutamate receptor; Dros, Drosophila; ND, not<br />

mentioned<br />

They are mild and non-denaturing as <strong>the</strong>y break lipid-lipid and lipid-protein interactions ra<strong>the</strong>r<br />

than protein-protein interactions (Seddon, Curnow et al. 2004). This allows membrane proteins to<br />

be solubilized without affecting <strong>the</strong> structural features <strong>of</strong> <strong>the</strong> protein (Sarramegna, Muller et al.<br />

2005). Ionic detergents contain ei<strong>the</strong>r a head group with a net charge that can be cationic (e.g.<br />

cetyl-trimethyl-ammonium bromide) or anionic (e.g. SDS), toge<strong>the</strong>r with a hydrophobic chain or a<br />

polar and an apolar face (e.g. sodium cholate). Among ionic detergents, SDS is effective in<br />

solubilizing membrane proteins but is also known to denature <strong>the</strong> protein (Sarramegna, Muller et<br />

al. 2005). However, some membrane proteins can be re-natured from SDS by transferring <strong>the</strong><br />

protein to ano<strong>the</strong>r detergent or lipid environment (Paulsen, Finkenzeller et al. 1993; Booth, Flitsch<br />

et al. 1995; Lau and Bowie 1997). Prior to a purification step, a suitable condition for solubilization<br />

is required for achieving maximum solubilization efficiency and stability <strong>of</strong> functional receptors in<br />

solution (Tucker and Grisshammer 1996; Warne, Chirnside et al. 2003). The detergent <strong>of</strong> choice<br />

must be made on a trial-and-error basis for each individual protein. The main criteria are <strong>the</strong><br />

compatibility <strong>of</strong> <strong>the</strong> chosen detergent with <strong>the</strong> purification steps and <strong>the</strong> necessity to preserve <strong>the</strong><br />

functionality <strong>of</strong> <strong>the</strong> receptor during <strong>the</strong> solubilization and purification steps (Sarramegna, Muller et<br />

al. 2005). The efficiency <strong>of</strong> detergents can be quantified by determining <strong>the</strong> receptor protein<br />

content in <strong>the</strong> solubilized sample before and after a centrifugation step at 100,000 x g. Radioactive<br />

40 | P a g e


GPCRs expression and purification<br />

ligands have been used to bind to <strong>the</strong> receptor and enable <strong>the</strong> quantification (Mazina, Strader et al.<br />

1996; de Jong, Grunewald et al. 2004). However, this technique detects only <strong>the</strong> functional fraction<br />

<strong>of</strong> <strong>the</strong> solubilized receptors, since in <strong>the</strong> presence <strong>of</strong> detergents, numerous proteins tend to lose<br />

<strong>the</strong>ir ability to bind ligands (Sarramegna, Muller et al. 2005). Ano<strong>the</strong>r important aspect in choosing<br />

detergents is <strong>the</strong> ease <strong>of</strong> removal <strong>of</strong> <strong>the</strong> detergents during a reconstitution step. In this case,<br />

detergents with a high critical micellar concentration (CMC) will be easier to eliminate<br />

(Sarramegna, Muller et al. 2005). However, detergents with low CMC can also be removed<br />

efficiently over a short period <strong>of</strong> time using polystyrene beads (Levy, Bluzat et al. 1990; Sansuk,<br />

Balog et al. 2008). Detergents with a low CMC such as digitonin, CHAPS and DM are used generally<br />

for GPCRs solubilization, since <strong>the</strong>y give a good efficiency <strong>of</strong> solubilization and a better stability <strong>of</strong><br />

protein (le Maire, Champeil et al. 2000). Some GPCRs require <strong>the</strong> use <strong>of</strong> several detergents; for<br />

example, neurotensin receptors were solubilized with an efficiency <strong>of</strong> 85% when using a mix <strong>of</strong><br />

detergents <strong>of</strong> CHAPS, DM and Cholesterol hemisuccinate (Tucker and Grisshammer 1996).<br />

The Purification step<br />

There are several different chromatographic methods which can be used to purify membrane<br />

proteins based on <strong>the</strong>ir properties. The most common protein purification technique for GPCRs is<br />

<strong>the</strong> affinity chromatography called Immobilized metal affinity chromatography (IMAC)<br />

(Janssen, Bovee-Geurts et al. 1995; Ratnala, Swarts et al. 2004; Sarramegna, Muller et al. 2005;<br />

Sansuk, Balog et al. 2008). The IMAC purification relies on <strong>the</strong> specific binding <strong>of</strong> <strong>the</strong> targeted<br />

protein to an immobilized phase. In order to do this, <strong>the</strong> targeted protein should be produced as a<br />

fusion protein <strong>of</strong> N- and/or C-term to an affinity tag. Among numerous numbers <strong>of</strong> affinity tags, <strong>the</strong><br />

histidine tag is <strong>the</strong> most commonly used tag (Arnau, Lauritzen et al. 2006). The histidine tag is very<br />

short and <strong>the</strong>refore, does not alter <strong>the</strong> solubility properties and rarely affect <strong>the</strong> characteristic <strong>of</strong><br />

<strong>the</strong> targeted protein (Graslund, Nordlund et al. 2008). During <strong>the</strong> purification step, <strong>the</strong> imidazole<br />

rings in <strong>the</strong> histidine tag bind to <strong>the</strong> divalent metal ions attached to an immobile phase, and enable<br />

<strong>the</strong> separation <strong>of</strong> <strong>the</strong> targeted protein from <strong>the</strong> o<strong>the</strong>r proteins or impurities by applying several<br />

washing steps. The targeted protein can be eluted from <strong>the</strong> immobilized phase by using a gradient<br />

<strong>of</strong> imidazole or by decreasing <strong>the</strong> pH <strong>of</strong> <strong>the</strong> buffer. In general, IMAC purification is performed ei<strong>the</strong>r<br />

in phosphate buffer (pH 8) or HEPES buffer (pH 7.5-8.0). An ionic strength, which maintains <strong>the</strong><br />

solubility and stability <strong>of</strong> <strong>the</strong> proteins, should be equivalent to 300-500 mM NaCl. Adding glycerol<br />

(~10%) in <strong>the</strong> buffer also aids <strong>the</strong> solubility and stability <strong>of</strong> <strong>the</strong> proteins. Reducing reagents such as<br />

41 | P a g e


Chapter 2<br />

dithiothreitol (DTT) should also be included in <strong>the</strong> buffer to maintain a reducing environment and<br />

finally, a small amount <strong>of</strong> imidazole in <strong>the</strong> buffer helps reducing <strong>the</strong> binding <strong>of</strong> less histidine-rich<br />

proteins to <strong>the</strong> IMAC column (Graslund, Nordlund et al. 2008).<br />

Table 3<br />

General properties <strong>of</strong> a range <strong>of</strong> common detergents<br />

Detergent type CMC (1) Comments<br />

Non-ionic detergents<br />

Triton X-100<br />

n-octyl-glucoside<br />

n-dodecyl-β-D-maltoside<br />

(DM)<br />

Ionic detergents<br />

Sodium dodecyl sulfate (SDS)<br />

Sodium cholate<br />

42 | P a g e<br />

0.24<br />

14.5<br />

0.16<br />

3.5<br />

13.0<br />

A mild detergent that is usually non-denaturing. Give a strong<br />

absorbance at 280 nM. Difficult to remove by dialysis owing to<br />

very low CMC. However, can be removed using BioBeads.<br />

Expensive to use for initial solubilization but good for<br />

reconstitution step owing to its high CMC.<br />

A mild detergent that can stabilize protein that is sensitive in<br />

solubilized form. It has a longer alkyl chain length than n-octyl<br />

glucoside, resulting in a lower CMC making this difficult to be<br />

removed.<br />

Denaturing to most membrane proteins. Commonly used in PAGE<br />

analysis.<br />

A non-denaturing detergent, but can disrupt protein-protein<br />

interactions in oligomeric complexes. Not compatible with ionexchange<br />

column chromatography.<br />

Zwitterionic detergents<br />

CHAPS<br />

4.0 Electrically neutral detergent with similar structure to Sodium<br />

cholate, but compatible with ion-exchange column<br />

chromatography.<br />

(1) CMC; Critical Micelle Concentration; <strong>the</strong> concentration <strong>of</strong> detergents/surfactants at which micelles<br />

spontaneously start forming.<br />

To date, <strong>the</strong>re is a broad array <strong>of</strong> resins used in <strong>the</strong> IMAC system such as TALON Metal Affinity<br />

(Clontech), <strong>the</strong> QIAexpressionist (QIAGEN) and HisTrap HP (GE Healthcare). Besides batch<br />

purification, <strong>the</strong> purification step can also be integrated by high-performance liquid<br />

chromatography (HPLC) system, such as ÄKTA from GE Healthcare. The IMAC purification has<br />

several advantages such as <strong>the</strong> strong and specific binding between <strong>the</strong> protein and <strong>the</strong><br />

immobilized phase, <strong>the</strong> mild elution conditions and <strong>the</strong> ability to control selectivity (by including<br />

low concentrations <strong>of</strong> imidazole in <strong>the</strong> chromatographic buffers). Fur<strong>the</strong>rmore, <strong>the</strong> purity <strong>of</strong> <strong>the</strong><br />

protein obtained from <strong>the</strong> IMAC can be easily optimized by controlling <strong>the</strong> ratio <strong>of</strong> recombinant<br />

protein to <strong>the</strong> column size (Graslund, Nordlund et al. 2008). Ano<strong>the</strong>r affinity chromatography<br />

technique is <strong>the</strong> Lectin affinity chromatography, which separates <strong>the</strong> protein based on <strong>the</strong>


GPCRs expression and purification<br />

affinity <strong>of</strong> proteoglycans for lectin. Only glycosylated protein is bound to <strong>the</strong> lectin resins and eluted<br />

with monosaccharides, α-D-methyl mannoside, α-D-methyl glycoside, or N-acetyl glucosamine<br />

(Cummings and Kornfeld 1982). There are different kinds <strong>of</strong> lectins available for <strong>the</strong> purification <strong>of</strong><br />

glycoproteins. Concanavalin A (ConA) or lentil lectin binds with a very high affinity to proteins<br />

presenting N-glycosylated amino acids with high-mannosidic carbohydrates. Wheat germ<br />

agglutinin (WGA) lectin selectively interacts with proteins containing complex N-glycans. The<br />

human β2AR expressed in Sf9 cells bind to ConA resin, but does not interact with a WGA resin,<br />

showing that <strong>the</strong> receptor mainly contains high-mannosidic sugars (Reilander, Boege et al. 1991).<br />

The rat M3 receptor expressed in Sf9 cells was able to bind to both WGA and ConA, demonstrating<br />

heterogeneity in glycosylation (Cummings and Kornfeld 1982), which underlines a limitation <strong>of</strong> this<br />

chromatographic method. In Biotin-Streptavidin affinity chromatography, <strong>the</strong> protein is fused<br />

with a 15-24 amino acid sequence called bio-tag. The tagged protein is biotinylated by biotin-<br />

protein ligase and subsequently purified using <strong>the</strong> affinity <strong>of</strong> biotin for avidin or Streptavidin<br />

(Sarramegn, Muller et al. 2006). A successful example is <strong>the</strong> purification <strong>of</strong> <strong>the</strong> NTS1 neurotensin<br />

receptor by Tucker and Grisshammer. The receptor protein is C-terminally fused with <strong>the</strong> Bio-tag<br />

and <strong>the</strong> protein is purified on an avidin column. The final yield is 38% with a remarkable 60-fold<br />

purification (Tucker and Grisshammer 1996). Besides <strong>the</strong>se affinity chromatography techniques,<br />

Ionic Exchange Chromatography is also a useful technique for protein purification. In this<br />

method, <strong>the</strong> protein can be separated on cation or anion exchange columns depending on <strong>the</strong>ir<br />

isoelectric point and <strong>the</strong> pH <strong>of</strong> <strong>the</strong> buffer (Sarramegn, Muller et al. 2006). Most <strong>of</strong> <strong>the</strong> protein<br />

purifications are done on anion exchange columns because proteins are generally negatively<br />

charged at physiological pH values (pH 6-8). The proteins are absorbed to <strong>the</strong> column, made from<br />

beads <strong>of</strong> inactive material such as cellulose or dextran, and <strong>the</strong>n eluted from <strong>the</strong> column by<br />

adjusting <strong>the</strong> salt concentration or changing <strong>of</strong> <strong>the</strong> pH. The proteins are eluted according to <strong>the</strong><br />

tenacity with which <strong>the</strong>y bind. Those that are <strong>the</strong> most highly charged at that pH will bind <strong>the</strong><br />

tightest. An example <strong>of</strong> this method is <strong>the</strong> purification <strong>of</strong> several GPCRs such as β2AR and<br />

muscarinic receptors by Parker and co-workers (Parker, Kameyama et al. 1991). Ion-exchange<br />

chromatography can also be performed as an additional purification step, if <strong>the</strong> target protein is<br />

contaminated with o<strong>the</strong>r proteins. For example, different phosphorylated states <strong>of</strong> protein can<br />

sometimes be resolved using ion-exchange chromatography (Bullock, Debreczeni et al. 2005).<br />

Since proteins can occur in many different sizes; we can <strong>the</strong>refore use this property for <strong>the</strong><br />

purification purpose. The separation based on <strong>the</strong> protein size is called Gel filtration<br />

43 | P a g e


Chapter 2<br />

chromatography or Size exclusion chromatography. This type <strong>of</strong> purification is more generic<br />

and can be performed in any buffer conditions. This technique has been used to resolve <strong>the</strong><br />

oligomerization state <strong>of</strong> <strong>the</strong> target protein (Graslund, Nordlund et al. 2008) or it is <strong>of</strong>ten used as a<br />

complement to o<strong>the</strong>r methods (Parker, Kameyama et al. 1991; Mazina, Strader et al. 1996; Tucker<br />

and Grisshammer 1996; Christ<strong>of</strong>fers, Li et al. 2003; Christ<strong>of</strong>fers, Li et al. 2005). Finally, after <strong>the</strong><br />

protein purification step, samples should be subjected to SDS-PAGE analysis. The purity and <strong>the</strong><br />

size <strong>of</strong> protein can be investigated after staining <strong>the</strong> SDS gel with a dye such as Coomassie Brilliant<br />

blue (Bradford 1976). To fur<strong>the</strong>r quantify <strong>the</strong> amount and concentration, The purified proteins can<br />

be measured at UV A280 (Graslund, Nordlund et al. 2008) or by a Bradford protein determination<br />

assay (Bradford 1976) to fur<strong>the</strong>r quantify <strong>the</strong> amount and concentration. Proper storage <strong>of</strong> <strong>the</strong><br />

purified protein is ano<strong>the</strong>r important aspect. The purified protein should be frozen in liquid<br />

nitrogen and stored at -80°C. It may be necessary to store <strong>the</strong> proteins in small aliquots to avoid<br />

damaging freeze-thaw cycles. If <strong>the</strong> proteins are found to be inactivated after one freeze-thaw cycle,<br />

<strong>the</strong>y can be stored at 4°C for short periods <strong>of</strong> time or at -20°C in high concentration <strong>of</strong> glycerol, or<br />

as an ammonium sulfate suspension (Graslund, Nordlund et al. 2008).<br />

The reconstitution step<br />

An ideal situation for <strong>the</strong> reconstitution <strong>of</strong> membrane protein such as GPCRs is to have an<br />

environment that more closely resembles <strong>the</strong> natural lipid bilayer. One <strong>of</strong> <strong>the</strong> strategies is <strong>the</strong><br />

incorporation <strong>of</strong> <strong>the</strong> GPCRs into liposomes. Liposomes can be made from a variety <strong>of</strong> lipids, such as<br />

phosphatidylserine, phosphatidylcholine, cholesterol, E.coli lipids or asolectin (Katzen, Peterson et<br />

al. 2009). Reconstitution <strong>of</strong> protein into such liposomes can be achieved by <strong>the</strong> addition <strong>of</strong> <strong>the</strong><br />

appropriate lipid-detergent micelles into <strong>the</strong> detergent-solubilized GPCRs, followed by <strong>the</strong> removal<br />

<strong>of</strong> detergent from <strong>the</strong> homogenous mixed solution (Gwyn 1994; Remigy, Caujolle-Bert et al. 2003).<br />

The detergent is dialyzed out over several hours, resulting in <strong>the</strong> formation <strong>of</strong> phospholipid vesicles<br />

(Gwyn 1994). However, this method requires a long dialysis process and cannot be used with low<br />

CMC detergents such as sodium cholate, and CHAPS (Chattopadhyay and London 1984). The<br />

dilution <strong>of</strong> <strong>the</strong> ternary mixture is a simpler variation <strong>of</strong> <strong>the</strong> dialysis method and has been shown to<br />

produce reconstituted membrane protein efficiently (Almog, Kushnir et al. 1986; Dolder, Engel et al.<br />

1996; Remigy, Caujolle-Bert et al. 2003). Yet, this method is again only possible when using high<br />

CMC detergents and it does not allow complete detergent removal (Rigaud, Mosser et al. 1997).<br />

Finally, a hydrophobic adsorption <strong>of</strong> <strong>the</strong> detergent to polystyrene beads (BioBeads) can be used to<br />

remove low CMC detergents (Akiba, Toyoshima et al. 1996; Rigaud, Mosser et al. 1997; Sansuk,<br />

44 | P a g e


GPCRs expression and purification<br />

Balog et al. 2008). The drawbacks <strong>of</strong> this method are its (too) fast detergent removal and <strong>the</strong><br />

adsorption <strong>of</strong> protein or lipid onto beads (Remigy, Caujolle-Bert et al. 2003). However, <strong>the</strong><br />

reconstitution process may not always be straightforward. If <strong>the</strong> detergent binds more strongly to<br />

<strong>the</strong> protein than to <strong>the</strong> lipids, <strong>the</strong> protein will be in its soluble form. Therefore, a fur<strong>the</strong>r detergent<br />

removal will precipitate <strong>the</strong> protein unless it is able to integrate into preformed lipid vesicles. In<br />

addition to this, <strong>the</strong> protein will also aggregate if it releases <strong>the</strong> detergent more readily than <strong>the</strong><br />

lipid does. Thus, an optimized reconstitution process should be carefully developed to have <strong>the</strong><br />

mixture consists mainly <strong>of</strong> ternary detergent-protein-lipid micelles (Remigy, Caujolle-Bert et al.<br />

2003). The success <strong>of</strong> <strong>the</strong> reconstitution process can be monitored by dynamic light scattering<br />

(Engel, Hoenger et al. 1992; Dolder, Engel et al. 1996), turbidimetry (Almog, Kushnir et al. 1986;<br />

Ollivon, Eidelman et al. 1988), and cryo-electron microscopy (Lambert, Levy et al. 1998). The most<br />

common method for detergent removal is dialysis (Jap, Walian et al. 1991; Kuhlbrandt, Wang et al.<br />

1994). For an optimized reconstitution process, <strong>the</strong> lipid-protein ratio (LPR) should be in <strong>the</strong> range<br />

<strong>of</strong> 0.2–2, and <strong>the</strong> protein concentration should not be less than 1 mg/ml (Remigy, Caujolle-Bert et<br />

al. 2003). Moreover, <strong>the</strong> choice <strong>of</strong> <strong>the</strong> detergent is critical. It must preserve <strong>the</strong> native state <strong>of</strong> <strong>the</strong><br />

protein during solubilization and purification, and possess <strong>the</strong> correctly balanced binding affinities<br />

to <strong>the</strong> protein and <strong>the</strong> lipid. In addition to this, <strong>the</strong> measurement <strong>of</strong> <strong>the</strong> detergent concentration<br />

prior to <strong>the</strong> reconstitution process is necessary since excess <strong>of</strong> detergent may destabilize<br />

membrane protein complexes, leading to a loss <strong>of</strong> <strong>the</strong> ternary structure (Remigy, Caujolle-Bert et al.<br />

2003). Fur<strong>the</strong>rmore, <strong>the</strong> selection <strong>of</strong> lipids is also <strong>of</strong> importance. In general, proteins exhibit<br />

optimal activity in artificial membranes with a lipid composition that most closely reflects that <strong>of</strong><br />

<strong>the</strong>ir native membranes (Gwyn 1994). For example <strong>the</strong> recovery <strong>of</strong> a D. melanogastor metabotropic<br />

receptor was dependent on <strong>the</strong> presence <strong>of</strong> ergosterol, <strong>the</strong> sterol originally found in Drosophila<br />

membrane (Gimpl and Fahrenholz 2002). It has also been shown that <strong>the</strong> kinetic <strong>of</strong> β2AR<br />

conformational changes can be improved by careful selection <strong>of</strong> lipids used for artificial membrane<br />

(Yao, Parnot et al. 2006). All <strong>of</strong> <strong>the</strong>se studies emphasize <strong>the</strong> importance <strong>of</strong> <strong>the</strong> reconstitution step to<br />

provide <strong>the</strong> most closed-to-native environment for purified GPCRs. Recently a new platform <strong>of</strong><br />

membrane environment has been developed. This membrane platform is so called <strong>the</strong> recombinant<br />

High Density Lipoprotein (rHDL) membrane or “nanodisc”. The nanodiscs are nanometer-sized,<br />

discoidal particles comprising amphipathic helical proteins that wrap <strong>the</strong>mselves around <strong>the</strong> planar<br />

circumference <strong>of</strong> a lipid bilayer (Borch and Hamann 2009; Katzen, Peterson et al. 2009). In contrary<br />

to <strong>the</strong> liposomes, nanodiscs are mono-disperse, highly stable and could provide a better<br />

representation <strong>of</strong> <strong>the</strong> phase transition behavior <strong>of</strong> membrane proteins such as GPCRs (Borch and<br />

45 | P a g e


Chapter 2<br />

Hamann 2009; Cappuccio, Hinz et al. 2009). Up to now, <strong>the</strong>re are several membrane proteins<br />

including <strong>the</strong> β2AR (Leitz, Bayburt et al. 2006) that have been successfully incorporated into such<br />

membrane discs (for fur<strong>the</strong>r review see ref (Borch and Hamann 2009)).<br />

46 | P a g e<br />

High resolution structures <strong>of</strong> GPCRs<br />

One <strong>of</strong> <strong>the</strong> most challenging tasks in membrane protein structural studies is to obtain highly<br />

diffracting crystals <strong>of</strong> GPCRs. Since, most <strong>of</strong> <strong>the</strong> GPCRs are expressed, except rhodopsin, at low<br />

levels in native tissues (Alkhalfioui 2009; Rosenbaum, Rasmussen et al. 2009; Topiol and Sabio<br />

2009). Therefore, a suitable recombinant expression system must be developed to generate<br />

natively folded membrane proteins. As shown in Table 4, only Sf9 insect cells (Rasmussen, Choi et<br />

al. 2007; Hanson, Cherezov et al. 2008; Jaakola, Griffith et al. 2008; Scheerer, Park et al. 2008;<br />

Warne, Serrano-Vega et al. 2008) and COS-7 mammalian cells (Standfuss, Xie et al. 2007) have<br />

produced enough purified GPCRs for structure determination. For bovine rhodopsin, a high level <strong>of</strong><br />

expression in native rod-cell disc membranes directly allows purification from a natural source<br />

(Palczewski, Kumasaka et al. 2000; Murakami and Kouyama 2008; Park, Scheerer et al. 2008;<br />

Scheerer, Park et al. 2008; Shimamura, Hiraki et al. 2008). Ano<strong>the</strong>r challenge in GPCRs<br />

crystallography is to overcome <strong>the</strong> protein stability problems. GPCRs have generally poor <strong>the</strong>rmal<br />

stability and <strong>the</strong>ir extra-membranous loops regions are prone to proteolysis (Topiol and Sabio<br />

2009). Various methods have been used to enhance <strong>the</strong> <strong>the</strong>rmal stability <strong>of</strong> <strong>the</strong> GPCRs, including <strong>the</strong><br />

use <strong>of</strong> ligands to stabilize <strong>the</strong> receptor (Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et al.<br />

2007; Hanson, Cherezov et al. 2008; Jaakola, Griffith et al. 2008; Warne, Serrano-Vega et al. 2008);<br />

<strong>the</strong> use <strong>of</strong> stabilizing mutants (Cherezov, Rosenbaum et al. 2007; Hanson, Cherezov et al. 2008;<br />

Warne, Serrano-Vega et al. 2008); <strong>the</strong> addition <strong>of</strong> lipids during purification and crystallization<br />

(Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007; Jaakola, Griffith et al. 2008), and<br />

having a high salt concentration (Jaakola, Griffith et al. 2008). Fur<strong>the</strong>rmore, various methods have<br />

also been used to enhance <strong>the</strong> proteolytic stability, including <strong>the</strong> truncation <strong>of</strong> disordered regions<br />

(Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007; Jaakola, Griffith et al. 2008; Warne,<br />

Serrano-Vega et al. 2008), <strong>the</strong> fusion <strong>of</strong> a stable protein at <strong>the</strong> ICL3 (Cherezov, Rosenbaum et al.<br />

2007; Jaakola, Griffith et al. 2008), and <strong>the</strong> complex formation with an antibody Fab fragment<br />

(Rasmussen, Choi et al. 2007). The crystallization <strong>of</strong> dark-state rhodopsin structures did not require<br />

such modifications due to <strong>the</strong>ir high stability (Palczewski, Kumasaka et al. 2000; Salom, Lodowski<br />

et al. 2006; Standfuss, Xie et al. 2007; Murakami and Kouyama 2008; Shimamura, Hiraki et al.<br />

2008). However, to obtain <strong>the</strong> active state opsin crystal (Park, Scheerer et al. 2008; Scheerer, Park


GPCRs expression and purification<br />

et al. 2008), <strong>the</strong> protein needed to be crystallized at low pH, since this condition is known to<br />

stabilize <strong>the</strong> active conformation <strong>of</strong> opsin (Topiol and Sabio 2009). After obtaining large quantities<br />

<strong>of</strong> purified and stable GPCRs, an effort in coaxing GPCRs into crystals is fur<strong>the</strong>r needed. For<br />

detergent-solubilized membrane protein such as GPCRs, <strong>the</strong> absence <strong>of</strong> exposed polar surface area<br />

outside <strong>the</strong> micelle can be a major limitation to crystallization (Alkhalfioui 2009; Rosenbaum,<br />

Rasmussen et al. 2009). Using <strong>of</strong> <strong>the</strong> antibody Fab complex approach [26] or <strong>the</strong> fusion <strong>of</strong> T4<br />

lysozyme strategy (Cherezov, Rosenbaum et al. 2007; Jaakola, Griffith et al. 2008) provided has<br />

provided sufficient polar surface in obtaining <strong>the</strong> crystals <strong>of</strong> adrenergic and adenosine receptors.<br />

All <strong>the</strong>se approaches give a promising sign that it is possible to obtain o<strong>the</strong>r similar engineered<br />

GPCR structures. However, when using <strong>the</strong>se alternative strategies, it is important to characterize<br />

<strong>the</strong> engineered GPCRs for <strong>the</strong>ir native-like pharmacology and biophysical properties (Rosenbaum,<br />

Rasmussen et al. 2009). The high resolution X-ray structures <strong>of</strong> GPCRs will enhance our<br />

understanding <strong>of</strong> <strong>the</strong> mode <strong>of</strong> action <strong>of</strong> agonists, inverse agonist and antagonists, and <strong>the</strong>reby<br />

accelerated drug discovery efforts aimed toward this class <strong>of</strong> membrane protein.<br />

Recombinant expression <strong>of</strong> <strong>the</strong> H1R protein<br />

Like <strong>the</strong> majority <strong>of</strong> GPCRs, <strong>the</strong> H1R proteins are found in small quantities in native membranes<br />

such as cat small intestinal muscle (Takagi and Uchida 1970), calf thymocytes (Osband and<br />

McCaffrey 1979), guinea pig brain (Toll and Snyder 1982), and rat liver (Fukui, Wang et al. 1988).<br />

In order to obtain a sufficient amount <strong>of</strong> <strong>the</strong> H1R protein for structural studies such as<br />

crystallography or NMR, it is necessary to express <strong>the</strong> recombinant protein in expression hosts. Up<br />

to date <strong>the</strong>re are reports showing <strong>the</strong> high expression level <strong>of</strong> H1R proteins in insect cells (Ratnala,<br />

Swarts et al. 2004; Akermoun, Koglin et al. 2005), yeast P. pastoris (Magnin, Fiez-Vandal et al. 2009)<br />

and <strong>the</strong> in vitro cell free expression system (Kamonchanok, Balog et al. 2008). Ratnala and co-<br />

workers reported in 2004 <strong>the</strong> functional expression <strong>of</strong> <strong>the</strong> H1R in insect Sf9 cells with at least 1000-<br />

fold higher than in native tissues (Ratnala, Swarts et al. 2004). The H1R protein was fur<strong>the</strong>r<br />

extracted by dodecyl-maltoside purified using affinity chromatography (IMAC) and reconstituted<br />

into artificial membrane proteins made from asolectin. In <strong>the</strong> end, <strong>the</strong>y were able to obtain up to 4-<br />

7 mg <strong>of</strong> <strong>the</strong> purified H1R protein from 1 Liter culture (Ratnala, Swarts et al. 2004). The milligram<br />

quantities <strong>of</strong> purified H1R have been subsequently used in <strong>the</strong>ir structural studies with NMR<br />

(Ratnala, Kiihne et al. 2007).<br />

47 | P a g e


Chapter 2<br />

Table 4<br />

Crystal structures <strong>of</strong> GPCRs<br />

GPCRs Species PDB Å Modifications In complex with Expression<br />

system<br />

bovine 1F88 (1) Rhodopsin<br />

2.8 None 11-cis-retinal Native<br />

Rhodopsin<br />

48 | P a g e<br />

bovine 2I35 (2)<br />

2I36 (2)<br />

2I37 (2)<br />

3.8<br />

4.1<br />

4.1<br />

Deprotonated Rhodopsin 11-cis-retinal Native<br />

bovine 2J4Y (3) 3.4 Mutation at N2C/D282C and<br />

N15D<br />

squid 2ZIY (4)<br />

2Z73 (5)<br />

3.7<br />

2.5<br />

Deletion <strong>of</strong> 75 residues at<br />

<strong>the</strong> C-term<br />

11-cis-retinal COS-7<br />

cells<br />

11-cis-retinal Native<br />

bovine 3CAP (6) 2.9 None None Native<br />

bovine 3DQB (7) 3.2 Contain syn<strong>the</strong>tic Gα at <strong>the</strong><br />

C-term<br />

None Native<br />

β2AR human 2RH1 (8) 2.4 Deletion <strong>of</strong> 48 residues at<br />

<strong>the</strong> C-term<br />

β2AR human 2R4R (9)<br />

2R4S (9)<br />

3.4<br />

3.4<br />

-Deletion <strong>of</strong> 48 residues at<br />

<strong>the</strong> C-term<br />

-Mutation at N187E<br />

-Fusion with T4-lysozyme in<br />

place <strong>of</strong> ICL3<br />

β2AR human 3D4S (10) 2.8 -Deletion <strong>of</strong> 65 residues at<br />

<strong>the</strong> C-term<br />

-Fusion with T4-lysozyme in<br />

place <strong>of</strong> ICL3<br />

-Mutation at E122W<br />

β1AR Turkey 2VT4 (11) 2.7 -Deletion <strong>of</strong> 30 residues at<br />

<strong>the</strong> N-term<br />

-Deletion <strong>of</strong> 28 residues in<br />

IC3<br />

-Deletion <strong>of</strong> 100 residues at<br />

<strong>the</strong> C-term<br />

-Various mutations to<br />

increase <strong>the</strong>rmal stability<br />

A2A<br />

Receptor<br />

Human 3EML (12) 2.6 -Deletion <strong>of</strong> 96 residues at<br />

<strong>the</strong> C-term<br />

-Fusion with T4-lysozyme in<br />

place <strong>of</strong> ICL3<br />

- Carazolol<br />

- ICL3 is bound<br />

with Antibody<br />

Fab fragment<br />

Insect<br />

cells<br />

Carazolol Insect<br />

cells<br />

Timolol<br />

Cholesterol<br />

Insect<br />

cells<br />

Cyanopindolol Insect<br />

cells<br />

ZM241385 Insect<br />

cells<br />

References: (1) Palczewski, Kumasaka et al. 2000, (2) Salom, Lodowski et al. 2006, (3) Standfuss, Xie et al. 2007,<br />

(4) Shimamura, Hiraki et al. 2008, (5) Murakami and Kouyama 2008, (6) Park, Scheerer et al. 2008, (7) Scheerer,<br />

Park et al. 2008, (8) Cherezov, Rosenbaum et al. 2007, (9) Rasmussen, Choi et al. 2007, (10) Hanson, Cherezov<br />

et al. 2008, (11) Warne, Serrano-Vega et al. 2008, (12) Jaakola, Griffith et al. 2008


GPCRs expression and purification<br />

In 2005, Akermoun and co-workers (Akermoun, Koglin et al. 2005) reported <strong>the</strong> characterization <strong>of</strong><br />

16 GPCRs, including <strong>the</strong> H1R, in various type <strong>of</strong> insect cells such as Sf9, Sf21 and High five. The<br />

highest expression level <strong>of</strong> <strong>the</strong> H1R is observed in Sf21 insect cells after 72h <strong>of</strong> infection.The<br />

expressed H1R protein has <strong>the</strong> Bmax <strong>of</strong> 85 pmol/mg protein when binding to [ 3H]mepyramine<br />

(Akermoun, Koglin et al. 2005). Magnin and co-workers have developed a rapid and generic<br />

purification process <strong>of</strong> various GPCRs expressed in P. pastoris. This purification relies on <strong>the</strong><br />

presence <strong>of</strong> a C-terminal biotinylation domain in <strong>the</strong> recombinant proteins, which binds with a high<br />

affinity to streptavidin-agarose beads. After <strong>the</strong> washing step, <strong>the</strong> specific tobacco etch virus (TEV)<br />

protease was added to remove <strong>the</strong> C-terminal tag and <strong>the</strong> native protein is released from <strong>the</strong><br />

agarose beads. Using this method, <strong>the</strong>y were able to obtain around 0.5 mg <strong>of</strong> <strong>the</strong> purified H1R from<br />

a liter culture <strong>of</strong> P. pastoris (Magnin, Fiez-Vandal et al. 2009). Recently, we also reported <strong>the</strong> use <strong>of</strong><br />

an in vitro cell free expression system to express <strong>the</strong> H1R protein (Kamonchanok, Balog et al. 2008).<br />

The cell free technology has several advantages as compared with <strong>the</strong> technology based on living<br />

organisms. These are <strong>the</strong> speed and directness <strong>of</strong> all procedures, <strong>the</strong> absence <strong>of</strong> constraints from<br />

living cells and <strong>the</strong> wide possibilities <strong>of</strong> product modifications. Using <strong>the</strong> cell free technology we<br />

were able to rapidly produce <strong>the</strong> sufficient amount <strong>of</strong> <strong>the</strong> H1R protein for <strong>the</strong> mass spectrometric<br />

analysis. These results will be fur<strong>the</strong>r discussed in <strong>the</strong> Chapter 3.<br />

49 | P a g e


CHAPTER<br />

Mass spectrometric analysis <strong>of</strong> <strong>the</strong> H1R:<br />

glycosylation and binding <strong>of</strong> an<br />

irreversible H1R antagonist<br />

Kamonchanok Sansuk 1, Crina Balog 2, Anne van der Does 2, Raymond<br />

Booth 3 , Wim de Grip 4 , Andre Deelder 2 , Remko Bakker 1 , Paul<br />

Hensbergen 2 , Rob Leurs 1<br />

1 Leiden/<strong>Amsterdam</strong> Center for Drug Research (LACDR), <strong>Vrije</strong> <strong>Universiteit</strong> <strong>Amsterdam</strong>,<br />

Department <strong>of</strong> Medicinal Chemistry, De Boelelaan 1083, 1081HV, <strong>Amsterdam</strong>, <strong>the</strong> Ne<strong>the</strong>rlands<br />

2 Biomolecular Mass Spectrometry Unit, Leiden University Medical Center (LUMC), Department<br />

<strong>of</strong> Parasitology, P.O. Box 9600, 2300 RC, Leiden, <strong>the</strong> Ne<strong>the</strong>rlands<br />

3 Department <strong>of</strong> Medicinal Chemistry, College <strong>of</strong> Pharmacy, University <strong>of</strong> Florida, Gainesville,<br />

Florida 32610-0485<br />

4 Department <strong>of</strong> Biophysical Organic Chemistry, LIC/LACDR, Gorlaeus Laboratories, Leiden<br />

University, P.O. Box 9502, 2300 RA, Leiden, <strong>the</strong> Ne<strong>the</strong>rlands<br />

5 Current address: Boehringer Ingelheim Pharma GmbH&Co. KG, Biberach, Germany<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

(Part <strong>of</strong> this research is published in: J Proteome Res, 2008, 7(2), 621-9)<br />

[Pick <strong>the</strong> date]


Chapter 4<br />

72 | P a g e<br />

Introduction<br />

G protein-coupled receptors (GPCRs) are <strong>the</strong> largest family <strong>of</strong> proteins in <strong>the</strong> human genome. They<br />

are membrane proteins with seven transmembrane (TM) domains that mediate responses to a<br />

broad range <strong>of</strong> stimuli such as light, taste, odor, hormones and neurotransmitters. GPCRs have been<br />

implicated in numerous human disorders and diseases, thus making <strong>the</strong>m attractive targets for<br />

drug discovery. It is estimated that nearly half <strong>of</strong> all modern drugs regulate GPCR activity in some<br />

way (Jacoby, Bouhelal et al. 2006; Schlyer and Horuk 2006; Lefkowitz 2007; Warne, Serrano-Vega<br />

et al. 2008). Despite <strong>the</strong>ir success as drug targets, <strong>the</strong> structural insight at <strong>the</strong> atomic level has long<br />

been limited to <strong>the</strong> crystal <strong>of</strong> rhodopsin (in its inactive state) (Palczewski, Kumasaka et al. 2000).<br />

Recently, three new GPCR proteins (Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007;<br />

Jaakola 2008; Warne, Serrano-Vega et al. 2008) and <strong>the</strong> activated state <strong>of</strong> bovine rhodopsin have<br />

been successfully crystallized (Park, Scheerer et al. 2008). These crystals provide significant new<br />

insight in GPCRs structures and <strong>the</strong>ir activation. However, similar information is still needed for<br />

o<strong>the</strong>r GPCRs that have not been successfully studied by X-ray crystallography yet. To date, <strong>the</strong>re are<br />

a variety <strong>of</strong> alternative techniques to detect receptor-ligand interactions. Mass spectrometry (MS)<br />

analysis has gained wide acceptance for <strong>the</strong> characterization <strong>of</strong> proteins. Proteins can be rapidly<br />

analyzed in <strong>the</strong> intact form or after <strong>the</strong> digestion with selected enzymes. Detection <strong>of</strong> <strong>the</strong> intact<br />

protein provides molecular mass information and <strong>the</strong> extent <strong>of</strong> protein heterogeneity. The analysis<br />

<strong>of</strong> proteolytic peptides allows confirmation <strong>of</strong> <strong>the</strong> amino acid sequence and a thorough<br />

investigation <strong>of</strong> any post-translational modification. With its high sensitivity, toge<strong>the</strong>r with<br />

continuous improvements in technology and computer s<strong>of</strong>tware, MS analysis has become not only a<br />

key player in protein sequencing but also a tool for <strong>the</strong> characterization <strong>of</strong> higher order protein<br />

structures, non-covalent complexes and receptor-ligand binding interactions (Sinz 2007). Up till<br />

now, <strong>the</strong>re are several GPCRs, such as rhodopsin (Ball, Oatis et al. 1998; Ablonczy, Kono et al. 2001;<br />

Kraft, Mills et al. 2001; Ablonczy, Crouch et al. 2005), <strong>the</strong> tachykinin NK-1 receptor (Alves, Sachon<br />

et al. 2007), <strong>the</strong> cannabinoid CB2 receptor (Zvonok, Yaddanapudi et al. 2007), <strong>the</strong> neurotensin NTS1<br />

receptor (Ho, White et al. 2008) and <strong>the</strong> histamine H1R (Sansuk, Balog et al. 2008) that had <strong>the</strong>ir<br />

protein sequences determined (almost) completely by MS analysis. We are particularly interested<br />

in <strong>the</strong> histamine H1R, one <strong>of</strong> <strong>the</strong> prototypic members <strong>of</strong> <strong>the</strong> family A GPCRs, which has been used as<br />

a model system for our GPCR production, purification and MS analysis (Sansuk, Balog et al. 2008).<br />

Using <strong>the</strong> developed protocol as described in Chapter 3, we have identified for <strong>the</strong> first time an<br />

asparagine residue in <strong>the</strong> H1R protein that is subjected to N-linked glycosylation. Fur<strong>the</strong>rmore, we


MS analysis <strong>of</strong> <strong>the</strong> H1R: glycosylation and binding <strong>of</strong> an irreversible H1R antagonist<br />

also showed preliminary data <strong>of</strong> <strong>the</strong> study <strong>of</strong> ligand-H1R binding interactions using<br />

phenoxybenzamine, an irreversible H1R antagonist.<br />

Reagents and Materials<br />

Experimental Procedures<br />

Standard laboratory chemicals, including phenoxybenzamine, were purchased from Sigma<br />

Chemical Co. (St. Louis, MO). Insect-Xpress insect medium, penicillin, and streptomycin were<br />

purchased from Bio Whittaker. Ni-NTA Agarose resin, NuPAGE Novex Bis-Tris Gel, and SimplyBlue<br />

Safestain were purchased from Invitrogen.<br />

Site-directed mutagenesis <strong>of</strong> <strong>the</strong> H1R<br />

The single point mutant human histamine H1R (N5Q H1R mutant) was created using a PCR-based<br />

mutagenesis approach (Cormack 2001). The constructs were sub-cloned into <strong>the</strong> expression vector<br />

pcDEF3 and verified by DNA sequencing.<br />

Cell culture and transfection<br />

COS-7 African green monkey kidney cells were maintained at 37 °C in a humidified 5% CO2/95% air<br />

atmosphere in Dulbecco’s Modified Eagle’s Medium (DMEM) containing 50 IU/ml penicillin, 50<br />

µg/ml streptomycin and 5% (v/v) fetal bovine serum. COS-7 cells were transiently transfected<br />

using <strong>the</strong> DEAE-dextran method as previously described (Bakker, Schoonus et al. 2001).<br />

Reporter-gene assay<br />

Reporter-gene assays were performed essentially as described previously (Bakker, Schoonus et al.<br />

2001). In brief, cells transiently co-transfected with pNF-κB-Luc (125 µg/1.10 7 cells) and pcDEF3<br />

encoding WT or N5Q H1R mutant (25 µg/1.10 7 cells) were seeded in 96 well white plates (Costar)<br />

in serum free culture medium and incubated with drugs. After 48 hrs, cells were assayed for<br />

luminescence by aspiration <strong>of</strong> <strong>the</strong> medium and <strong>the</strong> addition <strong>of</strong> 25 µl/well luciferase assay reagent<br />

(0.83 mM ATP. 0.83 mM d-luciferin, 18.7 mM MgCl2, 0.78 µM Na2H2P2O7, 38.9 mM Tris (pH 7.8).<br />

0.39% (v/v) glycerol, 0.03% (v/v) Triton X-100 and 2.6 µM dithiothreitol). After 30 min,<br />

luminescence was measured for 3 sec/well in a Victor plate reader (Wallac).<br />

73 | P a g e


Chapter 4<br />

Baculovirus Expression System and <strong>the</strong> treatment <strong>of</strong> recombinantly expressed H1R protein with<br />

phenoxybenzamine<br />

Spodoptera frugiperda ovarian cells (Sf9) were cultured as monolayer at 27 ºC in tissues culture<br />

flasks in a serum-free insect medium supplemented with 50 units/mL penicillin and 50 mg/mL<br />

streptomycin. Under this conditions, <strong>the</strong> cell doubling time was typically 20-24 h. Cells were<br />

infected with pBac-H1His10 virus at a multiplicity <strong>of</strong> infection (MOI) <strong>of</strong> 0.1. Infected Sf9 cells were<br />

maintained in a serum-free insect medium supplemented with 5 μM Leupeptin, 50 units/mL<br />

penicillin and 50 mg/mL streptomycin and harvested at 5 days post infection (dpi). For <strong>the</strong> pre-<br />

incubation with phenoxybenzamine, <strong>the</strong> infected Sf9 cells were incubated with phenoxybenzamine<br />

at <strong>the</strong> concentration <strong>of</strong> 30-90 nM in Bis-Tris propane buffer (20 mM Bis-Tris propane, 1 M NaCl, 1<br />

mM Histidine, 5 mM β-mercaptoethanol, and 5 µM Leupeptin, pH 7.8). After 30 min incubation, <strong>the</strong><br />

Bis-Tris propane buffer containing phenoxybenzamine was removed and <strong>the</strong> infected Sf9 cells were<br />

washed with fresh Bis-Tris propane buffer for at least 3 times.<br />

Radioligand Binding Assays<br />

The infected Sf9 cells or <strong>the</strong> transfected COS-7 cells were harvested and homogenized in <strong>the</strong> H1R<br />

binding buffer (50 mM Na2/K Phosphate buffer, pH 7.4). The membranes were <strong>the</strong>n incubated with<br />

5 nM [ 3H] mepyramine (specific activity 30 Ci/mmol) (Perkin-Elmer) in a final volume <strong>of</strong> 200 µL for<br />

30 min at 25 °C. Nonspecific binding was determined by incubation with 1 µM <strong>of</strong> <strong>the</strong> selective H1R<br />

antagonist mianserin. Non-bound radioactivity was removed by filtration through Whatman GF/C<br />

filters that had been treated with 0.3% polyethyleneimine. Filters were washed 3–4 times with ice-<br />

cold H1R binding buffer, and radioactivity retained on <strong>the</strong> filters was measured by liquid<br />

scintillation counting. Binding data was evaluated by a nonlinear, least-square, curve-fitting<br />

procedure using GraphPad Prism (GraphPad S<strong>of</strong>tware, Inc., San Diego, CA).<br />

Protein purification <strong>of</strong> <strong>the</strong> H1R protein<br />

The infected Sf9 cell membranes was prepared based on <strong>the</strong> protocol according to Ratnala and co-<br />

workers (Ratnala, Swarts et al. 2004). The membranes were subsequently solubilized in Bis-Tris<br />

propane buffer, containing 2% (w/v) n-dodecyl-β- D-maltoside (DDM), under gentle rotation at 4 °C<br />

for 4 h. The supernatant was separated by ultracentrifugation at 80,000 g for 20 min at 4 °C and<br />

incubated with 2 mL Ni-NTA agarose resin, pre-equilibrated with Bis-Tris propane buffer. After<br />

overnight incubation at 4 °C, Ni-NTA agarose resin was settled by gravity, and <strong>the</strong> unbound lysate<br />

74 | P a g e


MS analysis <strong>of</strong> <strong>the</strong> H1R: glycosylation and binding <strong>of</strong> an irreversible H1R antagonist<br />

was aspirated. The resin was washed with Bis-Tris propane buffer, containing 0.2% (w/v) DDM,<br />

and imidazole (concentration 20 and 100 mM). The bound H1R proteins were eluted from <strong>the</strong> Ni-<br />

NTA agarose resin with Bis-Tris propane buffer, containing 0.2% (w/v) DDM and 200 mM<br />

imidazole. Purified H1R protein was analyzed by gel electrophoresis, followed by protein staining or<br />

Western blotting.<br />

Gel Electrophoresis and Western Blotting<br />

Protein samples were mixed with SDS-PAGE sample buffer (2% (w/v) sodium dodecyl sulfate, 0.04<br />

M dithiothreitol, and 0.015% bromophenol blue in 0.5 M Tris-HCl, final pH 6.8 to a final volume <strong>of</strong><br />

25–50 µL. Samples were run on a NuPAGE Novex Bis-Tris Gel at 200 V constant, according to <strong>the</strong><br />

manufacturer’s protocol. For immunodetection, proteins were blotted onto a nitrocellulose<br />

membrane (1 h at 100 V) in ice-cold blot buffer (25 mM Tris and 0.2 M glycine in 20% methanol).<br />

Blots were subsequently immunoassayed for <strong>the</strong> presence <strong>of</strong> His-tagged receptor, using anti-His<br />

antibody (1:2000 dilution) (Qiagen). Protein staining was performed using Coomassie G-250 stain<br />

(SimplyBlue Safestain).<br />

In-Gel Tryptic Digestion and Deglycosylation<br />

Protein bands were excised from a NuPAGE Novex Bis-Tris gel, cut into small pieces, and washed<br />

with 25 mM NH 4HCO 3 followed by dehydration with 100% acetonitrile (ACN) for 10 min. For<br />

reduction and alkylation, dried gel particles were first incubated with 10 mM dithiothreitol for 30<br />

min at 56 °C. Following dehydration with ACN, gel plugs were subsequently incubated in 55 mM<br />

iodoacetamide for 20 min at room temperature. After washing with 25 mM NH 4HCO 3 and<br />

dehydration with 100% ACN, <strong>the</strong> gel particles were completely dried in a centrifugal vacuum<br />

concentrator (Eppendorf, Hamburg, Germany). For in-gel digestion using RapiGest TM SF, dried gel<br />

particles were re-swollen in a 1% RapiGest TM SF solution (in 25 mM NH 4HCO 3) followed by 10 min<br />

incubation at 37 °C. Excess solution was removed, and <strong>the</strong> gel particles were dried in a centrifugal<br />

vacuum concentrator. Dried gel particles were subsequently re-swollen on ice and incubated with<br />

10-15 μL <strong>of</strong> a trypsin solution (12.5 ng trypsin/μL 25 mM NH4HCO3) for 10 min at 37 °C. Following<br />

tryptic digestion, <strong>the</strong> overlaying digestion-solution was collected. After that, 0.5% trifluoroacetic<br />

acid (TFA) (final concentration) was added to <strong>the</strong> digestion solution followed by incubation for 40<br />

min at 37 °C. The sample was centrifuged at 20,000 g, for 10 min. Finally, <strong>the</strong> resulting supernatant<br />

was collected, and <strong>the</strong> pellet was extracted two times with 7.5 µL <strong>of</strong> 70% isopropanol. For<br />

subsequent deglycosylation, <strong>the</strong> supernatant was adjusted to pH 8 using ammonium bicarbonate<br />

75 | P a g e


Chapter 4<br />

and incubated with 5 Units <strong>of</strong> PNGase F (Roche Diagnostics, Mannheim, Germany) overnight at 37<br />

°C.<br />

MALDI-ToF MS<br />

MALDI-ToF(-ToF) analyses were performed on an Ultraflex II mass spectrometer (Bruker Daltonics,<br />

Bremen, Germany) using dihydroxybenzoic acid (5 mg/ml in 50%ACN/0.1% TFA) as a matrix. The<br />

mass spectrometer was used in <strong>the</strong> reflector ion mode. For fragment ion analysis in <strong>the</strong> tandem<br />

time-<strong>of</strong>-flight (ToF/ToF) mode, precursors were selected in a timed ion gate. Fragment ions<br />

generated by laser-induced decomposition <strong>of</strong> <strong>the</strong> precursor were fur<strong>the</strong>r accelerated in <strong>the</strong> LIFT<br />

cell, and <strong>the</strong>ir masses were analyzed after <strong>the</strong> ion reflector passage.<br />

Nano-LC-Iontrap MS<br />

The samples containing <strong>the</strong> digested H1R in 70% isopropanol were concentrated to approximately<br />

1–2 µL in a centrifugal vacuum concentrator. Subsequently, 7 µL <strong>of</strong> 70% ACN/0.1%TFA was added,<br />

and <strong>the</strong> sample was sonified for 10 min. Next, 0.1% TFA was added until a concentration <strong>of</strong> 20%<br />

ACN/0.1% TFA was reached. This was <strong>the</strong>n shortly sonified (2 min) just prior to injection onto <strong>the</strong><br />

nanoLC system (Ultimate, Dionex-LC Packings, The Ne<strong>the</strong>rlands). The sample was separated on a<br />

C18 pepmap100 column (75 µm i.d., 15 cm long; Dionex-LC packings, <strong>Amsterdam</strong>, The Ne<strong>the</strong>rlands)<br />

coupled with an electrospray ionization (ESI)-iontrap mass spectrometer (HCTultra, Bruker<br />

Daltonics, Bremen). The samples were injected onto <strong>the</strong> same system as described previously in <strong>the</strong><br />

chapter 3 <strong>of</strong> this <strong>the</strong>sis (see experimental procedures section) <strong>the</strong> samples were eluted using a<br />

linear gradient <strong>of</strong> 0% solvent B (95% ACN/0.1% formic acid) up to 60% solvent B in 60 min.<br />

MS/MS analysis was performed in a data-dependent manner using <strong>the</strong> five most intense peaks in an<br />

MS spectrum.<br />

Analysis <strong>of</strong> H1R deglycosylation<br />

76 | P a g e<br />

Results and Discussion<br />

The analysis <strong>of</strong> GPCR glycosylation have relied mainly on indirect methods such as apparent<br />

molecular weight shifts after Western blotting <strong>of</strong> tunicamycin-treated cells or PNGase- treated<br />

protein extracts (Kim, Zhang et al. 2005; Zhang, Kim et al. 2007). Moreover, site-directed<br />

mutagenesis <strong>of</strong> asparagines within N-glycosylation consensus sites has yielded valuable<br />

information, not only about <strong>the</strong> exact localization <strong>of</strong> <strong>the</strong> glycan moiety, but also on <strong>the</strong> functional


MS analysis <strong>of</strong> <strong>the</strong> H1R: glycosylation and binding <strong>of</strong> an irreversible H1R antagonist<br />

consequences (Rands, Candelore et al. 1990; Kohno, Wada et al. 2002). Because <strong>of</strong> <strong>the</strong> common<br />

difficulties in proteomic analysis <strong>of</strong> GPCRs, studies applying MS techniques for <strong>the</strong> direct<br />

assessment <strong>of</strong> <strong>the</strong> glycosylation sites in GPCRs are restricted to rhodopsin (Duffin, Lange et al.<br />

1993; Barnidge, Dratz et al. 1999; Abdulaev and Ridge 2000; Nakagawa, Miyamoto et al. 2001;<br />

Ablonczy, Goletz et al. 2002). As described in Chapter 3, <strong>the</strong> H1R protein obtained from <strong>the</strong><br />

baculovirus/insect cell expression system appeared as a smear (see Chapter 3, Figure 1) possibly<br />

reflecting H1R glycosylation. Within <strong>the</strong> amino terminal domain, <strong>the</strong> H1R contains 2 sequences<br />

(N5SS and N18KT) that fit with <strong>the</strong> consensus sequence for N-glycosylation (Asn-Xaa-Ser/Thr, Xaa<br />

being any amino acid except proline). It is hi<strong>the</strong>rto unknown whe<strong>the</strong>r <strong>the</strong>se sites are actually<br />

glycosylated. To address this question, we first analyzed a tryptic digest <strong>of</strong> H1R before and after<br />

treatment with PNGase F, using MALDI-ToF MS. PNGase F cleaves asparagine N-linked<br />

oligosaccharides at <strong>the</strong> β-aspartyl-glycosylamine linkage, converting asparagine into aspartic acid.<br />

Within <strong>the</strong> PNGase F-treated sample, we identified a peptide at m/ z 1405.7 which was absent in<br />

<strong>the</strong> untreated sample (Figure 1A). MALDI-ToF-ToF analysis revealed that it corresponds to <strong>the</strong><br />

acetylated N-terminal tryptic peptide <strong>of</strong> <strong>the</strong> H1R containing an aspartic acid at position 5. This<br />

aspartic acid corresponds to <strong>the</strong> converted asparagine (N) within <strong>the</strong> consensus sequence Asn-Ser-<br />

Ser after PNGase F releases, identifying this N5 as a glycan attachment site. However, in <strong>the</strong>se<br />

experiments we did not observe glycosylation at N18. Although this probably reflects site-specific<br />

glycosylation at N5, we cannot formally exclude <strong>the</strong> possibility <strong>of</strong> protein glycosylation at N18 as<br />

well. N5 was not fully glycosylated since we also identified <strong>the</strong> non-glycosylated counterpart <strong>of</strong> this<br />

peptide. To analyze <strong>the</strong> N5 glycosylation in more detail, we specifically searched for glycopeptides<br />

within our LC-iontrap MS data through <strong>the</strong> analysis <strong>of</strong> <strong>the</strong> glycopeptide marker ion at m/ z 366<br />

(hexose1HexNAc1 (Wuhrer, Catalina et al. 2007)) in our MS/MS spectra. Two major groups <strong>of</strong><br />

glycopeptides corresponding to different glycan structures attached to <strong>the</strong> Asn 5-containing tryptic<br />

peptide were observed. The first contained a heterogeneous population <strong>of</strong> high mannose structures<br />

(exemplified in Figure 1B). The o<strong>the</strong>r major group consists <strong>of</strong> N-glycans with terminal GlcNAc-<br />

residues with or without fucosyl residues linked to <strong>the</strong> N, N′-diacetyl-chitobiose core (exemplified in<br />

Figure 1C). These glycan structures were predicted based on what is known from glycosylation in<br />

insect cells (Altmann, Staudacher et al. 1999). Since glycosylation in insect cells does not resemble<br />

mammalian type <strong>of</strong> glycosylation, future MS analysis <strong>of</strong> H1R protein purified from insect Sf9 cells,<br />

stably expressing mammalian glycosyltransferases (Mimic TM Sf9 insect cells), or purified from<br />

mammalian cells overexpressing <strong>the</strong> H1R should be performed to reveal <strong>the</strong> mammalian<br />

glycosylation pattern <strong>of</strong> <strong>the</strong> H1R protein.<br />

77 | P a g e


Chapter 4<br />

Figure 1. Identification <strong>of</strong> <strong>the</strong> H1R glycosylation<br />

Tryptic peptides from Sf9 cell produced H1R were deglycosylated using PNGase F and analysed by MALDI-<br />

ToF MS. (A) A peptide at m/z 1405.7 was selected for MALDI-ToF-ToF and identified as <strong>the</strong> N-terminal<br />

tryptic peptide from H1R containing as Asp residue at position 4 which corresponds to <strong>the</strong> Asn5 residue in<br />

H1R. (B, C) Glycopeptides containing Asn5 were fur<strong>the</strong>r analysed by LC-iontrap MS/MS. Red triangle, fucose;<br />

blue square, N-acetylglucosamine; green circle, mannose.<br />

78 | P a g e


MS analysis <strong>of</strong> <strong>the</strong> H1R: glycosylation and binding <strong>of</strong> an irreversible H1R antagonist<br />

Many GPCRs are glycosylated in vivo, and glycosylation has been shown to be implicated in proper<br />

GPCR expression, signaling, and trafficking (Walsh, Foley et al. 1998; Wheatley and Hawtin 1999;<br />

Zhang, Austin et al. 2001; Zvonok, Yaddanapudi et al. 2007). To fur<strong>the</strong>r evaluate whe<strong>the</strong>r <strong>the</strong><br />

glycosylation at N5 plays a role in <strong>the</strong> expression, signaling and trafficking <strong>of</strong> <strong>the</strong> H1R, we <strong>the</strong>refore<br />

mutated <strong>the</strong> asparagine into glutamine (N5Q H1R mutant). The N5Q H1R mutant was expressed in<br />

COS-7 cells and its binding property was fur<strong>the</strong>r evaluated with <strong>the</strong> radioligand [ 3H]mepyramine.<br />

As shown in Figure 2A, <strong>the</strong> N5Q H1R mutant was well expressed in COS-7 cells with <strong>the</strong> Bmax <strong>of</strong><br />

12.4±9.7 pmol/mg protein (N=3) and has an affinity towards <strong>the</strong> radioligand [ 3H]mepyramine with<br />

<strong>the</strong> KD <strong>of</strong> 0.9±0.1 nM (N=3), similar to that <strong>of</strong> <strong>the</strong> wild type H1R (Bmax = 14.8±4.6 pmol/mg protein,<br />

N=7 and KD =1.5±0.2, N=7). This result demonstrates that <strong>the</strong> N-glycosylation at Asparagine 5 is not<br />

essential for <strong>the</strong> H1R expression and <strong>the</strong> targeting <strong>of</strong> H1R protein to <strong>the</strong> plasma membrane. The N5Q<br />

H1R mutant was also tested for <strong>the</strong> <strong>the</strong> functionality using NFκB reportergene assay. Figure 2B<br />

shows that both <strong>the</strong> wild type and <strong>the</strong> mutant H1R still possess basal activity and could be<br />

stimulated upon induction with its ligand histamine (pEC50 <strong>of</strong> <strong>the</strong> wild type H1R: 6.5±0.1, N=5;<br />

pEC50 <strong>of</strong> <strong>the</strong> N5Q H1R mutant: 6.5±0.1, N=5 ). Fur<strong>the</strong>rmore, <strong>the</strong> inverse agonist mepyramine<br />

inhibited <strong>the</strong> basal activity <strong>of</strong> both <strong>the</strong> wild type and <strong>the</strong> N5Q H1R mutant (pIC50 <strong>of</strong> <strong>the</strong> wild type<br />

H1R: 7.9±0.1, N=3; pEC50 <strong>of</strong> <strong>the</strong> N5Q H1R mutant: 8.0±0.1, N=3).<br />

Figure 2. Characterization <strong>of</strong> <strong>the</strong> N5Q H1R mutant, compared to <strong>the</strong> wild type H1R<br />

(A) The wild type (�, dotted line) and <strong>the</strong> N5Q H1R mutant ( , straight line) were expressed in COS-7 cells<br />

and fu<strong>the</strong>r analyzed for <strong>the</strong>ir expression level and KD values by radioligand binding analysis using<br />

[ 3H]mepyramine. (B) The H1R constructs were expressed in COS-7 cells. Upon induction with histamine,<br />

both <strong>the</strong> wild type (�, dotted line) and <strong>the</strong> N5Q H1R mutant ( , straight line) show <strong>the</strong> increasing NFκB<br />

activation and <strong>the</strong> inverse agonist mepyramine was able to inhibit <strong>the</strong> basal response <strong>of</strong> both <strong>the</strong> wild type<br />

(�, dotted line) and <strong>the</strong> N5Q H1R mutant (�, straight line).<br />

79 | P a g e


Chapter 4<br />

These results suggest that <strong>the</strong> mutation <strong>of</strong> <strong>the</strong> H1R at <strong>the</strong> N-linked glycosylation site (N5) did not<br />

affect its functional coupling to Gq pathway. However, as mentioned earlier that <strong>the</strong> glycosylation <strong>of</strong><br />

<strong>the</strong> H1R is also occurred at <strong>the</strong> residue N18, it might be worthwide to characterize <strong>the</strong> H1R<br />

constructs with a single mutation at N18 and double mutations at N5 and N18 for future studies on<br />

<strong>the</strong> importance <strong>of</strong> N-linked glycosylation in <strong>the</strong> H1R.<br />

MS analysis <strong>of</strong> phenoxybenzamine treated H1R<br />

Phenoxybenzamine is known as a classical α-adrenergic receptor (α-AR) antagonist. However, <strong>the</strong><br />

clinical use <strong>of</strong> phenoxybenzamine has been limited due to its long lasting α-AR blockade. Besides<br />

being known as a blocking agent for α-AR, phenoxybenzamine is also capable <strong>of</strong> inhibiting H1R<br />

responses (Cook 1971; Cook 1988; Rappen-Cremer 1989; Takayanagi 1996).<br />

Figure 3. Chemical structure <strong>of</strong> phenoxybenzamine<br />

The structure <strong>of</strong> phenoxybenzamine and its reactive aziridinium intermediate in aqueous solution are<br />

shown. The figure is obtained from Frang, et al. 2001 (Frang, Cockcr<strong>of</strong>t et al. 2001).<br />

Cook and co-workers have determined <strong>the</strong> effects <strong>of</strong> phenoxybenzamine on histamine-induced<br />

contractions <strong>of</strong> isolated guinea pig ileum. They found that at a phenoxybenzamine concentration <strong>of</strong><br />

5.9 x 10 -7 M and an exposure time <strong>of</strong> 5 min, a shift <strong>of</strong> about a half log unit is observed without <strong>the</strong><br />

depression <strong>of</strong> <strong>the</strong> maximum response. When increasing <strong>the</strong> concentration, <strong>the</strong> maximum response<br />

is depressed, and at <strong>the</strong> concentration <strong>of</strong> 5 x 10 -5 M and higher, <strong>the</strong> response is totally abolished<br />

(Cook 1971). A similar inhibition pr<strong>of</strong>ile <strong>of</strong> phenoxybenzamine is also observed in <strong>the</strong> work <strong>of</strong><br />

Takayanaki and co-workers. A progressive (> 70 min) treatment <strong>of</strong> guinea pig ileum tissue with<br />

phenoxybenzamine at a concentration <strong>of</strong> 10 -6 M inhibits <strong>the</strong> response <strong>of</strong> H1R to histamine<br />

(Takayanagi 1996). Fur<strong>the</strong>rmore, it has also been shown that phenoxybenzamine could block <strong>the</strong><br />

response induced by a partial agonist 2-(2-pyridyl)-ethylamine (PEA) (Cook 1988; Rappen-Cremer<br />

80 | P a g e


MS analysis <strong>of</strong> <strong>the</strong> H1R: glycosylation and binding <strong>of</strong> an irreversible H1R antagonist<br />

1989). It is thought that in an aqueous solution, phenoxybenzamine forms an unstable aziridinium<br />

ion (Figure 3) which can alkylate side chains <strong>of</strong> amino acid residues, such as –SH, -OH, =NH, and –<br />

COOH, and resulting in an irreversible binding between <strong>the</strong> ligand and <strong>the</strong> target protein (Jenkinson<br />

1996). It has also been reported that <strong>of</strong> <strong>the</strong> susceptible amino acid residues, cysteine with <strong>the</strong> –SH<br />

side chain is <strong>the</strong> most reactive (Shulman-Roskes, Noe et al. 1998). Prior to <strong>the</strong> mapping <strong>of</strong><br />

phenoxybenzamine binding site on <strong>the</strong> H1R protein, we first tested whe<strong>the</strong>r phenoxybenzamine<br />

would bind irreversibly to <strong>the</strong> H1R. In this case, we treated <strong>the</strong> recombinantly expressed H1R from<br />

baculovirus/insect cells expression system with various concentration <strong>of</strong> phenoxybenzamine (30-<br />

90 nM) for 30 min. The Bmax level <strong>of</strong> <strong>the</strong> recombinantly expressed H1R, compares to that <strong>of</strong> non-<br />

phenoxybenzamine treatment, was measured afterwards by <strong>the</strong> radioligand binding analysis using<br />

[ 3H]mepyramine. As shown in Figure 4A, <strong>the</strong> alkylating effect <strong>of</strong> phenoxybenzamine on <strong>the</strong> H1R<br />

protein blocked <strong>the</strong> binding <strong>of</strong> [ 3H]mepyramine to <strong>the</strong> H1R by 53%, 77% and 98%, when treating<br />

<strong>the</strong> receptor at <strong>the</strong> phenoxybenzamine concentration <strong>of</strong> 30, 60 and 90 nM, respectively.<br />

Figure 4. Effect <strong>of</strong> <strong>the</strong> phenoxybenzamine treatment on <strong>the</strong> H1R protein<br />

(A) The H1R protein expressed on Sf9 cells were incubated in <strong>the</strong> absence and presence (30, 60 and 90 nM)<br />

<strong>of</strong> phenoxybenzamine (PB) for 30 min at 37 °C. (B, top) The H1R protein samples from purification process<br />

were analyzed by western blot using anti-His antibody. The protein bands were present at 55 kDa. (B,<br />

bottom) Bmax values <strong>of</strong> <strong>the</strong> H1R proteins were determined using Radioligand binding analysis; (1) H1R<br />

membrane, (2) Phenoxybenzamine treated H1R membrane, (3) Phenoxybenzamine treated H1R after<br />

solubilization, (4) Phenoxybenzamine treated H1R after purification.<br />

A similar alkylating effect <strong>of</strong> phenoxybenzamine was also observed on <strong>the</strong> α-AR (Frang, Cockcr<strong>of</strong>t et<br />

al. 2001). Besides showing <strong>the</strong> irreversible binding effect <strong>of</strong> phenoxybenzamine on <strong>the</strong> H1R, <strong>the</strong>se<br />

81 | P a g e


Chapter 4<br />

data also suggest that phenoxybenzamine may have at least some part <strong>of</strong> its binding site in common<br />

with H1R antagonists such as mepyramine (Cook 1988). In order to map <strong>the</strong> amino acid residue that<br />

may be involved in <strong>the</strong> interaction <strong>of</strong> phenoxybenzamine and <strong>the</strong> H1R protein using MS analysis, we<br />

fur<strong>the</strong>r purified <strong>the</strong> phenoxybenzamine-treated H1R membrane (incubation time 30 min with <strong>the</strong><br />

phenoxybenzamine concentration at 90 nM) using IMAC purification system as previously<br />

described in Chapter 3. Samples were collected at various stages <strong>of</strong> purification and <strong>the</strong>y were<br />

tested on <strong>the</strong> western blot and radioligand binding analysis. As shown in Figure 4B, we were able<br />

to detect <strong>the</strong> protein band representing <strong>the</strong> H1R protein (~ 55 kDa) at every stages <strong>of</strong> purification.<br />

On <strong>the</strong> o<strong>the</strong>r hands, when performing radioligand binding analysis on <strong>the</strong>se collected samples, we<br />

still observed <strong>the</strong> blockade <strong>of</strong> [ 3H]mepyramine binding on <strong>the</strong> H1R by phenoxybenzamine. These<br />

results ensure <strong>the</strong> presence <strong>of</strong> H1R proteins and <strong>the</strong> irreversibly binding <strong>of</strong> phenoxybenzamine on<br />

<strong>the</strong> H1R protein. The purified phenoxybenzamine-treated H1R protein were subjected to one<br />

dimensional gel electrophoresis and stained with Coomassie Brilliant blue. The H1R protein bands<br />

at 55 kDa were excised and analyzed by MALDI-ToF MS analysis. The experimentally determined<br />

peptide masses from <strong>the</strong> phenoxybenzamine-treated and <strong>the</strong> non-treated H1R samples matched<br />

well. We were also able to obtain a tryptic fragment from <strong>the</strong> phenoxybenzamine-treated H1R with<br />

a mass difference corresponding to phenoxybenzamine (Figure 5).<br />

Unexpectedly, <strong>the</strong> cysteine residue that was identified to be alkylated by phenoxybenzamine is at<br />

<strong>the</strong> position 302 (Cys302) <strong>of</strong> <strong>the</strong> third intracellular loop <strong>of</strong> <strong>the</strong> H1R. Based on <strong>the</strong> structural insight<br />

<strong>of</strong> <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs (see Chapter 1 for fur<strong>the</strong>r reading ), <strong>the</strong> binding pocket <strong>of</strong> <strong>the</strong> H1R<br />

is thought to reside within <strong>the</strong> TM domains, such as TM3, TM5 and TM6. Fur<strong>the</strong>rmore, it has been<br />

shown that <strong>the</strong> cysteine residue involved in <strong>the</strong> binding <strong>of</strong> phenoxybenzamine in <strong>the</strong> α-adrenergic<br />

receptor is located at position 3.36 <strong>of</strong> TM3. We, <strong>the</strong>refore, expected to map <strong>the</strong> interaction <strong>of</strong><br />

phenoxybenzamine with amino acid residues inside <strong>the</strong> transmembrane domain region.<br />

Interestingly, Espinoza-Fonseca and Trujillo-Ferrara have proposed in 2006 <strong>the</strong> binding <strong>of</strong> an<br />

allosteric modulator, KT5720, to <strong>the</strong> intracellular region <strong>of</strong> <strong>the</strong> muscarinic M1 receptors (Espinoza-<br />

Fonseca 2006) and 2 years later, Nicholls and co-workers have suggested <strong>the</strong> binding <strong>of</strong> <strong>the</strong> CXCR2<br />

antagonist SB332235 to <strong>the</strong> intracellular region <strong>of</strong> CXCR2 (Nicholls 2008). It is tempting to<br />

speculate that <strong>the</strong> H1R can also be modulated at <strong>the</strong> intracellular site, although it is also possible<br />

that <strong>the</strong> covalent interaction between <strong>the</strong> Cys302 <strong>of</strong> <strong>the</strong> H1R and phenoxybenzamine from this<br />

preliminary study is <strong>the</strong> result from a non-specific reaction, since <strong>the</strong> binding <strong>of</strong> phenoxybenzamine<br />

to non-specific sites has also been observed previously (Terner 1971). Never<strong>the</strong>less, <strong>the</strong> mapping <strong>of</strong><br />

82 | P a g e


MS analysis <strong>of</strong> <strong>the</strong> H1R: glycosylation and binding <strong>of</strong> an irreversible H1R antagonist<br />

a cysteine residue in <strong>the</strong> H1R protein that is reactive to phenoxybenzamine provides evidence that<br />

MS technology is definitely a promising proteomic tool in identifying amino acid residues involved<br />

in ligand-protein interactions. For future MS analysis, <strong>the</strong> protocol regarding phenoxybenzamine<br />

treatment and <strong>the</strong> H1R sample preparation can be adjusted to minimize potential non-specific<br />

reactions. Fur<strong>the</strong>rmore, it is also noteworthy to fur<strong>the</strong>r investigate <strong>the</strong> intracellular region <strong>of</strong> <strong>the</strong><br />

H1R as a possible binding site by mutation analysis toge<strong>the</strong>r with modeling studies.<br />

Figure 5. MALDI-ToF-MS analysis <strong>of</strong> <strong>the</strong> pre-treated H1R with phenoxybenzamine<br />

A tryptic fragment from phenoxybenzamine-treated H1R was recovered and selected for MS/MS based on<br />

<strong>the</strong> increase in <strong>the</strong> mass corresponding to phenoxybenzamine ( 268.2). The fragmentation pattern clearly<br />

confirms that this peptide contains a phenoxybenzamine, most likely conjugated to <strong>the</strong> cysteine residue.<br />

83 | P a g e


Chapter 4<br />

84 | P a g e<br />

Conclusion<br />

The work presented here provides a promising MS analysis protocol that allows <strong>the</strong><br />

characterization <strong>of</strong> post-translational modifications (N-Linked glycosylation) and ligand binding<br />

sites <strong>of</strong> <strong>the</strong> H1R. We succeeded for <strong>the</strong> first time in identifying a single amino acid residue (N5) in<br />

<strong>the</strong> H1R that is glycosylated upon expression in insect cells. Fur<strong>the</strong>rmore, we also showed our<br />

preliminary results in using <strong>the</strong> irreversible ligand phenoxybenzamine to identify direct ligand-H1R<br />

protein interaction. Taken toge<strong>the</strong>r, <strong>the</strong> use <strong>of</strong> affinity labels or o<strong>the</strong>r irreversible ligands toge<strong>the</strong>r<br />

with <strong>the</strong> MS technology will allow us to progress our studies on ligand-receptor interaction. The<br />

precise knowledge <strong>of</strong> receptor bound ligand structures would provide opportunities for drug<br />

discovery and aid <strong>the</strong> development <strong>of</strong> tailor-made medicines substantially.


CHAPTER<br />

A structural insight into <strong>the</strong> reorientation<br />

<strong>of</strong> transmembrane domains 3 and 5 during<br />

H1R activation<br />

Kamonchanok Sansuk 1 , Xavier Deupi 2 , Ivan Torrecillas 2 , Aldo Jongejan 1 ,<br />

Remko Bakker 1, 3 , Leonardo Pardo 2 , Rob Leurs 1<br />

1Leiden/<strong>Amsterdam</strong> Center for Drug Research (LACDR), <strong>Vrije</strong> <strong>Universiteit</strong> <strong>Amsterdam</strong>,<br />

Department <strong>of</strong> Medicinal Chemistry, De Boelelaan 1083, 1081HV, <strong>Amsterdam</strong>, <strong>the</strong> Ne<strong>the</strong>rlands<br />

2Laboratori de Medicina Computacional, Unitat de Bioestadistica, Facultat de Medicina,<br />

Universitat Autonoma de Barcelona, 08193 Bellaterra, Spain<br />

3Current address: Boehringer Ingelheim Pharma GmbH&Co. KG, Biberach, Germany<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

(This research is submitted to Journal <strong>of</strong> Molecular Pharmacology)<br />

[Pick <strong>the</strong> date]


Chapter 5<br />

86 | P a g e<br />

Introduction<br />

G protein-coupled receptors (GPCRs) transduce sensory signals <strong>of</strong> external origin such as photons,<br />

odors or pheromones and endogenous signals, including neurotransmitters like histamine,<br />

(neuro)peptides, proteases, glycoprotein hormones and ions, to <strong>the</strong> cytoplasmic side <strong>of</strong> <strong>the</strong> cell<br />

membrane (Kristiansen 2004). The mechanism by which binding <strong>of</strong> <strong>the</strong>se highly diverse chemical<br />

signals triggers a set <strong>of</strong> conformational rearrangements <strong>of</strong> <strong>the</strong> transmembrane (TM) segments near<br />

<strong>the</strong> G-protein binding domains remains largely unknown. NMR studies on <strong>the</strong> human histamine H1<br />

receptor (H1R), a rhodopsin family type GPCR, using 13C, 15N-labelled histamine suggested that <strong>the</strong><br />

transition <strong>of</strong> <strong>the</strong> H1R from an predominantly inactive state to an active state after agonist binding<br />

involves proton transfer, similar to <strong>the</strong> meta I/meta II transition equilibrium in rhodopsin (Ratnala<br />

et al. 2007). Comparison <strong>of</strong> <strong>the</strong> structure <strong>of</strong> inactive rhodopsin (Li, Edwards et al. 2004) with <strong>the</strong><br />

recent crystal structure <strong>of</strong> <strong>the</strong> ligand-free opsin (Park, Scheerer et al. 2008), which contains several<br />

distinctive features <strong>of</strong> <strong>the</strong> presumed active state, leads to <strong>the</strong> conclusion that during <strong>the</strong> process <strong>of</strong><br />

GPCR activation, among o<strong>the</strong>rs changes, TM3 rotates clockwise (viewed from <strong>the</strong> intracellular side),<br />

<strong>the</strong> intracellular part <strong>of</strong> TM6 tilts outwards by 6-7 Å, TM5 nears TM6, and R3.50 within <strong>the</strong><br />

(D/E)R(Y/W) motif in TM3 adopts an extended conformation pointing towards <strong>the</strong> protein core<br />

(Park, Scheerer et al. 2008). Notably, as revealed in <strong>the</strong> original publication <strong>of</strong> opsin, <strong>the</strong>se<br />

conformational changes disrupt <strong>the</strong> ionic interaction between R3.50 with negatively charged side<br />

chains at positions 3.49 and 6.30 (Smit, Vischer et al. 2007; Nygaard, Frimurer et al. 2009) and<br />

facilitates <strong>the</strong> interactions between <strong>the</strong> highly conserved Y5.58 and K5.66 in TM5 and R3.50 in TM3<br />

and E6.30 in TM6, respectively. This strongly suggests that both TM3 and TM5 play a central role in<br />

stabilizing <strong>the</strong> active state <strong>of</strong> GPCRs and, <strong>the</strong>refore, in <strong>the</strong> process <strong>of</strong> GPCR activation. Previously,<br />

we have reported that binding <strong>of</strong> histamine to <strong>the</strong> H1R initiates an interaction between TM3 and<br />

TM7. Specifically, we found that S3.36 interacts with N7.49, freeing W6.48, which is o<strong>the</strong>rwise fixed<br />

in <strong>the</strong> inactive formation by N7.49. The S3.36-N7.49 serves as <strong>the</strong> first step in a cascade <strong>of</strong> o<strong>the</strong>r<br />

transmembrane rearrangements that eventually lead to <strong>the</strong> receptor assuming <strong>the</strong> activated state,<br />

in which it is able to activate its G-protein (Jongejan, Bruyster et al. 2005). In this study we set out<br />

to fur<strong>the</strong>r our knowledge <strong>of</strong> <strong>the</strong> activation <strong>of</strong> <strong>the</strong> H1R by combining <strong>the</strong> latest structural insights in<br />

GPCR structure, computational modeling and site-directed mutagenesis <strong>of</strong> <strong>the</strong> H1R to study <strong>the</strong> role<br />

<strong>of</strong> <strong>the</strong> TAS(I/V) motif at positions 3.37-3.40 in TM3, highly conserved in amine receptors, in <strong>the</strong><br />

process <strong>of</strong> receptor activation. Our results suggest that T3.37 interacts with TM5 in <strong>the</strong> inactive<br />

state <strong>of</strong> <strong>the</strong> H1R, while S3.39 changes conformation upon receptor activation to achieve, toge<strong>the</strong>r<br />

with S3.36, new water-mediated inter-helical interactions with TM2 and TM7 that stabilize <strong>the</strong>


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

active state. In addition, we provide <strong>the</strong> first evidence that <strong>the</strong> hydrophobic side chain at position<br />

3.40, highly conserved in <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs, plays a key role in facilitating <strong>the</strong><br />

histamine-induced relocation <strong>of</strong> TM5. Mutation <strong>of</strong> I3.40 to ei<strong>the</strong>r Ala or Gly, i.e. removing <strong>the</strong> bulky<br />

side chain at this position, abolishes <strong>the</strong> constitutive activity <strong>of</strong> <strong>the</strong> histamine H1 receptor (H1R), <strong>the</strong><br />

effect <strong>of</strong> constitutive-activity increasing mutations, as well as <strong>the</strong> histamine-induced receptor<br />

activation.<br />

Materials<br />

Experimental procedures<br />

Gifts <strong>of</strong> mianserin hydrochloride (Schering-Plough, The Ne<strong>the</strong>rlands), pcDEF3 (Dr. J. Langer<br />

(Goldman, Cutrone et al. 1996)) and <strong>of</strong> <strong>the</strong> cDNA encoding <strong>the</strong> human histamine H1R (Dr. H. Fukui<br />

(Ohta, Hayashi et al. 1994)) are greatly acknowledged. pNF-κB-Luc was obtained from Stratagene<br />

(La Jolla, USA), ATP disodium salt, bovine serum albumin, chloroquine diphosphate, DEAE-dextran<br />

(chloride form), histamine dihydrochloride, mepyramine (pyrilamine maleate) and<br />

polyethylenemine were purchased from Sigma Chemical Company (USA). D-luciferin was obtained<br />

from Duchefa Biochemie BV (Haarlem, The Ne<strong>the</strong>rlands), glycerol from Riedel-de-Haen (Germany),<br />

and Triton X-100 from Fluka (Switzerland). Cell culture media, penicillin and streptomycin were<br />

obtained from Invitrogen Corporation (Paisley, UK). Fetal bovine serum was obtained from Intergro<br />

B.V. (Dieren, The Ne<strong>the</strong>rlands). Cell culture plastics were obtained from Corning Costar (NY, USA).<br />

[ 3H]mepyramine (30 Ci/mmol) was purchased-from ICN Biomedicals BV (Zoetermeer, The<br />

Ne<strong>the</strong>rlands)<br />

Cell culture and transfection<br />

COS-7 African green monkey kidney cells were maintained at 37 °C in a humidified 5% CO2/95% air<br />

atmosphere in Dulbecco’s Modified Eagle’s Medium (DMEM) containing 50 IU/ml penicillin, 50<br />

µg/ml streptomycin and 5% (v/v) fetal bovine serum. COS-7 cells were transiently transfected<br />

using <strong>the</strong> DEAE-dextran method as previously described (Bakker, Schoonus et al. 2001).<br />

Site-directed mutagenesis<br />

Single point mutant human histamine H1Rs: T112A (T3.37A), T112E (T3.37E), I115A (I3.40A),<br />

I115G (I3.40G), S114A (S3.39A), S114C (S3.39C) and S114T (S3.39T) and double mutant human<br />

histamine H1Rs: S3.36T/I3.40A and S3.36T/I3.40G were created using a PCR-based mutagenesis<br />

87 | P a g e


Chapter 5<br />

approach (Cormack 2001). All constructs were sub-cloned into <strong>the</strong> expression vector pcDEF3 and<br />

verified by DNA sequencing.<br />

Reporter-gene assay<br />

Reporter-gene assays were performed essentially as described previously (Bakker, Schoonus et al.<br />

2001). In brief, cells transiently co-transfected with pNF-κB-Luc (125 µg/1.10 7 cells) and pcDEF3<br />

encoding WT or mutant H1Rs (25 µg/1.10 7 cells) were seeded in 96 well white plates (Costar) in<br />

serum free culture medium and incubated with drugs. After 48 hrs, cells were assayed for<br />

luminescence by aspiration <strong>of</strong> <strong>the</strong> medium and <strong>the</strong> addition <strong>of</strong> 25 µl/well luciferase assay reagent<br />

(0.83 mM ATP. 0.83 mM d-luciferin, 18.7 mM MgCl2, 0.78 µM Na2H2P2O7, 38.9 mM Tris (pH 7.8).<br />

0.39% (v/v) glycerol, 0.03% (v/v) Triton X-100 and 2.6 µM dithiothreitol). After 30 min,<br />

luminescence was measured for 3 sec/well in a Victor plate reader (Wallac).<br />

Histamine H1R binding studies<br />

Histamine H1R binding studies were performed essentially as described previously (Bakker,<br />

Schoonus et al. 2001). In brief, transfected COS-7 cells used for radioligand binding studies were<br />

harvested after 48hrs and homogenized in ice-cold 50 mM Na2/K-phosphate buffer (pH 7.4)<br />

(binding buffer). The COS-7 cell homogenates were incubated for 30 min at 30 °C in binding buffer<br />

in 200 µl with 3 nM [ 3H]mepyramine. Non-specific binding was determined in <strong>the</strong> presence <strong>of</strong> 1 µM<br />

mianserin. The incubations were stopped by rapid dilution with 3 ml ice-cold binding buffer. The<br />

bound radioactivity was separated by filtration through Whatman GF/C filters that had been<br />

treated with 0.3% polyethyleneimine. Filters were washed twice with 3 ml binding buffer and<br />

radioactivity retained on <strong>the</strong> filters was measured by liquid scintillation counting.<br />

Analytical methods<br />

Protein concentrations were determined according to Bradford (Bradford 1976), using BSA as a<br />

standard. Binding and functional data were evaluated by a non-linear, least squares curve-fitting<br />

procedure using GraphPad Prism 4 ® (GraphPad S<strong>of</strong>tware, Inc., San Diego, CA).<br />

Computational model <strong>of</strong> <strong>the</strong> histamine H1R<br />

A model <strong>of</strong> <strong>the</strong> TM domains 1-7 <strong>of</strong> <strong>the</strong> histamine H1R was build by homology modeling techniques<br />

using <strong>the</strong> crystal structure <strong>of</strong> <strong>the</strong> β2AR (PDB code 2RH1) (Cherezov, Rosenbaum et al. 2007;<br />

88 | P a g e


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

Rasmussen, Choi et al. 2007) as template. Building this homology model was straightforward except<br />

for <strong>the</strong> highly conserved NPxxYxn=5,6F motif at <strong>the</strong> junction between TM7 and <strong>the</strong> intracellular helix<br />

8. Notably, this junction is one residue shorter in <strong>the</strong> β2AR (n=5) than in rhodopsin (n=6). As a<br />

result, Y7.53 is pointing towards TM2 in rhodopsin (Li, Edwards et al. 2004) and towards TM6 in<br />

<strong>the</strong> β2AR (Rasmussen, Choi et al. 2007). The fact that <strong>the</strong> histamine H1R contains six amino acids<br />

(n=6) between both aromatic side chains led us to model <strong>the</strong> junction between TM7 and helix 8 as<br />

in rhodopsin (PDB code 1GZM). SCWRL-3.0 (Canutescu, Shelenkov et al. 2003) was employed to<br />

add <strong>the</strong> side chains <strong>of</strong> <strong>the</strong> non-conserved residues, using a backbone-dependent rotamer library.<br />

Modeller 9v1 (Marti-Renom, Stuart et al. 2000) was used to add intracellular loops IL1-2 and<br />

extracellular loops EL1-3 using <strong>the</strong> structure <strong>of</strong> <strong>the</strong> β2AR as template. Internal water molecules 506,<br />

519, 528, 529, 532, 534, 537, 543, 546, and 548 that mediate a number <strong>of</strong> inter-helical interactions<br />

(Rasmussen, Choi et al. 2007) and are probably conserved in <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs<br />

(Pardo, Deupi et al. 2007) were explicitly included in <strong>the</strong> model.<br />

Distortion <strong>of</strong> TM5<br />

The program HELANAL (Bansal, Kumar et al. 2000) was used to calculate <strong>the</strong> residue-residue unit<br />

twist angle <strong>of</strong> TM5, a structural parameter that describes local helical uniformity. This parameter is<br />

interpreted as follows: an ideal α-helix, with approximately 3.6 residues per turn, has a twist angle<br />

<strong>of</strong> approximately 100º (360º/3.6); a closed helical segment, with 100º; whereas an open helical segment, with >3.6 residues per turn, possesses a twist<br />


Chapter 5<br />

receptor <strong>the</strong> constitutive activity decreases (right column). On <strong>the</strong> contrary, a mutation that<br />

decreases <strong>the</strong> strength <strong>of</strong> an intramolecular interaction (i.e., Ser to Ala, Asp to Asn, …) (Table 1A,<br />

bottom row) leads to an increase (left column) or a decrease (right column) <strong>of</strong> <strong>the</strong> constitutive<br />

activity depending whe<strong>the</strong>r <strong>the</strong> mutated residue is stabilizing <strong>the</strong> inactive or <strong>the</strong> active state <strong>of</strong> <strong>the</strong><br />

receptor, respectively. In addition, single point mutations that modify <strong>the</strong> side chain rotamer<br />

distribution, without modifying its capability to form intramolecular interactions, can help to<br />

elucidate <strong>the</strong> conformer involved in <strong>the</strong> stabilization <strong>of</strong> ei<strong>the</strong>r <strong>the</strong> inactive or active state <strong>of</strong> <strong>the</strong><br />

receptor (Table 1B).<br />

Analysis <strong>of</strong> structural databases shows that Ser, Thr, and Cys residues possess different side chain<br />

rotamer distribution when located in a �-helical structure (Ballesteros, Deupi et al. 2000). Ser can<br />

adopt ei<strong>the</strong>r <strong>the</strong> g+ (52% <strong>of</strong> <strong>the</strong> side chains), g- (20%), or t (28%) rotamer conformation; <strong>the</strong> β-<br />

branched side chain <strong>of</strong> Thr is limited to <strong>the</strong> g+ (85%) and g- (15%) because <strong>the</strong> t conformation is<br />

unfavorable due to <strong>the</strong> steric clash <strong>of</strong> <strong>the</strong> side chain methyl group with <strong>the</strong> backbone carbonyl at <strong>the</strong><br />

i-3 position; while Cys is restricted to <strong>the</strong> g+ (71%) or t (29%) conformation because <strong>of</strong> <strong>the</strong> steric<br />

clash between <strong>the</strong> Sγ atom and <strong>the</strong> carbonyl oxygen <strong>of</strong> residue i-3 in <strong>the</strong> g- conformation<br />

(McGregor, Islam et al. 1987). Thus, if mutation <strong>of</strong> Ser (able to adopt <strong>the</strong> g+, g- or t conformations)<br />

to Cys (g+ and t) (Table 1B, top row) and Thr (g+ and g-) (Table 1B, bottom row) increases, in both<br />

cases, <strong>the</strong> constitutive activity (left column), Ser adopts <strong>the</strong> g+ conformation in <strong>the</strong> active state, as<br />

both mutations lead to an increase <strong>of</strong> <strong>the</strong> relative population <strong>of</strong> <strong>the</strong> g+ conformation. Conversely, if<br />

<strong>the</strong>se mutations lead to a decrease <strong>of</strong> <strong>the</strong> constitutive activity (right column), <strong>the</strong> g+ conformation<br />

<strong>of</strong> Ser would stabilize <strong>the</strong> inactive conformation <strong>of</strong> <strong>the</strong> receptor.<br />

Numbering Scheme <strong>of</strong> GPCRs<br />

Residues are identified by <strong>the</strong> generic numbering scheme <strong>of</strong> Ballesteros & Weinstein (Ballesteros<br />

and Weinstein 1995) that allows easy comparison among residues in <strong>the</strong> 7TM segments <strong>of</strong> different<br />

receptors.<br />

Sequence analysis<br />

Family-specific analysis <strong>of</strong> GPCR sequences were performed using <strong>the</strong> GMoS program, freely<br />

available at http://lmc.uab.cat/gmos.<br />

90 | P a g e


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

Table 1 Correspondence between single point mutations, experimentally determined constitutive<br />

agonist-independent activity, and receptor structure<br />

Combination <strong>of</strong> structural bioinformatics with experimental results allows deducing <strong>the</strong><br />

structural/functional role <strong>of</strong> <strong>the</strong> mutated residues.<br />

STRUCTURAL BIOINFORMATICS<br />

A<br />

B<br />

Effect <strong>of</strong> <strong>the</strong><br />

mutation on <strong>the</strong><br />

strength <strong>of</strong> <strong>the</strong><br />

intramolecular<br />

interaction<br />

Effect <strong>of</strong> <strong>the</strong><br />

mutation on <strong>the</strong><br />

side chain<br />

conformation<br />

distribution<br />

increase strength<br />

(e.g. Ala → Ser)<br />

decrease strength<br />

(e.g. Ser → Ala)<br />

Ser → Cys<br />

g+/g-/t g+/t<br />

Ser → Thr<br />

g+/g-/t g+/g-<br />

EXPERIMENTAL RESULTS<br />

Constitutive activity (CA)<br />

increase CA decrease CA<br />

Side chain<br />

stabilizes <strong>the</strong><br />

ACTIVE<br />

conformation<br />

Side chain<br />

stabilizes <strong>the</strong><br />

INACTIVE<br />

conformation<br />

Ser in g+<br />

stabilizes <strong>the</strong><br />

ACTIVE<br />

conformation<br />

Side chain<br />

stabilizes <strong>the</strong><br />

INACTIVE<br />

conformation<br />

Side chain<br />

stabilizes <strong>the</strong><br />

ACTIVE<br />

conformation<br />

Ser in g+<br />

stabilizes <strong>the</strong><br />

INACTIVE<br />

conformation<br />

91 | P a g e


Chapter 5<br />

92 | P a g e<br />

Results<br />

A proline-induced unwinding in <strong>the</strong> structure <strong>of</strong> TM5 <strong>of</strong> Class A GPCRs.<br />

GPCRs, including <strong>the</strong> H1R, possess a highly conserved P5.50 in TM5. This conserved proline<br />

presents in 77% <strong>of</strong> <strong>the</strong> rhodopsin-like sequences (Mirzadegan, Benko et al. 2003), and absents only<br />

in melanocortin, glycoprotein hormone, lysosphingolipid, prostanoid, and cannabinoid receptors.<br />

Usually, in Pro-containing α-helices, <strong>the</strong> steric clash between <strong>the</strong> pyrrolidine ring <strong>of</strong> Pro and <strong>the</strong><br />

backbone carbonyl oxygen in <strong>the</strong> preceding turn induces a bend angle <strong>of</strong> approximately 20º in <strong>the</strong><br />

helical structure (Deupi, Olivella et al. 2004). In contrast, <strong>the</strong> structures <strong>of</strong> bovine (Li, Edwards et al.<br />

2004) and squid (Murakami and Kouyama 2008) rhodopsin, adenosine A2A receptor (Jaakola,<br />

Griffith et al. 2008), and <strong>the</strong> β1AR (Warne, Serrano-Vega et al. 2008) and β2AR (Cherezov,<br />

Rosenbaum et al. 2007; Rasmussen, Choi et al. 2007) have revealed a peculiar conformation <strong>of</strong> TM5<br />

in which <strong>the</strong> Pro-kink induced bend is decreased (Figure 1A) due to a local opening <strong>of</strong> <strong>the</strong> helix<br />

(>3.6 residues/turn, twist < 100º, see Experimental Procedures) at <strong>the</strong> 5.45-5.48 turn (Figure 1B)<br />

that partially removes <strong>the</strong> steric clash between <strong>the</strong> pyrrolidine ring <strong>of</strong> P 5.50 and <strong>the</strong> backbone<br />

carbonyl oxygen at position 5.46 (i-4 in Figure 1C and 1D). This local opening <strong>of</strong> TM5 (proline-<br />

unwinding, in contrast to proline-kink) also modifies <strong>the</strong> relative orientation <strong>of</strong> <strong>the</strong> side chains at <strong>the</strong><br />

extracellular side (Govaerts, Lefort et al. 2001; Deupi, Dolker et al. 2007), including residues<br />

involved in neurotransmitter binding such as 5.42, 5.43, and 5.46 (Shi, Liapakis et al. 2002). O<strong>the</strong>r<br />

membrane proteins also feature this type <strong>of</strong> helical wide turns (Riek, Finch et al. 2008).<br />

The role <strong>of</strong> TM3 in stabilizing <strong>the</strong> proline-unwinding <strong>of</strong> TM5<br />

The mechanism by which this unusual conformation <strong>of</strong> TM5 is stabilized remains unknown. In<br />

bovine rhodopsin, <strong>the</strong> backbone carbonyl oxygen at position 5.46 is stabilized by a hydrogen bond<br />

interaction with E3.37 (at <strong>the</strong> inter-atomic distance <strong>of</strong> 2.8Å) and a van der Waals interaction with<br />

L3.40 (3.2Å) (Figure 1C) (Li, Edwards et al. 2004; Deupi, Dolker et al. 2007). Adenosine A2A<br />

receptor preserves similar interactions through Q3.37 (3.2Å) and I3.40 (3.9Å) (Jaakola, Griffith et<br />

al. 2008). Interestingly, squid rhodopsin replaces E3.37 <strong>of</strong> bovine rhodopsin by Phe but presents a<br />

discrete water molecule linking <strong>the</strong> backbone carbonyls at positions 3.37 (3.0Å) and 5.46 (3.3Å)<br />

(Murakami and Kouyama 2008), while maintaining <strong>the</strong> van der Waals interaction between <strong>the</strong><br />

hydrophobic I3.40 and <strong>the</strong> carbonyl oxygen at position 5.46 (3.2Å). Notably, <strong>the</strong> recent X-ray<br />

structure <strong>of</strong> opsin (Park, Scheerer et al. 2008) shows that during <strong>the</strong> process <strong>of</strong> GPCR activation <strong>the</strong><br />

hydrogen bond interaction between E3.37 and <strong>the</strong> backbone carbonyl (4.9Å) is disrupted, as


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

previously suggested by NMR measurements <strong>of</strong> rhodopsin containing 13C-labeled histidine (Patel,<br />

Crocker et al. 2005), while L3.40 has moved slightly away (3.5Å) (Figure 1D).<br />

This effect correlates with <strong>the</strong> minor, but significant, clockwise rotation (viewed from <strong>the</strong><br />

intracellular side) <strong>of</strong> TM3 during <strong>the</strong> process <strong>of</strong> receptor activation (see Discussion). The release <strong>of</strong><br />

<strong>the</strong> strong packing <strong>of</strong> L3.40 with <strong>the</strong> 5.45-5.48 turn does not result in a change in <strong>the</strong> local opening<br />

<strong>of</strong> TM5 (Figure 1B), but in a localized decrease in <strong>the</strong> bend around P5.50 (Figure 1A). In <strong>the</strong> β1AR<br />

(Warne, Serrano-Vega et al. 2008) and β2AR (Cherezov, Rosenbaum et al. 2007; Rasmussen, Choi et<br />

al. 2007), I3.40 also stabilizes <strong>the</strong> carbonyl oxygen at position 5.46 (i-4 in Figure 1E), similarly to<br />

adenosine and opsin receptors, whereas, in this case, <strong>the</strong> shorter T3.37 (compared to E or Q)<br />

interacts with <strong>the</strong> side chain, ra<strong>the</strong>r than with <strong>the</strong> backbone carbonyl, <strong>of</strong> S5.46. To investigate <strong>the</strong><br />

role <strong>of</strong> <strong>the</strong> amino acids at positions 3.37 and 3.40 in GPCR activation we have created and studied<br />

several histamine H1 receptor (H1R) mutants. We have previously used <strong>the</strong> H1R as a model system<br />

for <strong>the</strong> study <strong>of</strong> Class A GPCR activation (Bakker, Jongejan et al. 2008). The H1R belongs to <strong>the</strong><br />

aminergic subfamily <strong>of</strong> Class A GPCRs, and like <strong>the</strong> β1AR and β2AR features T3.37 and I3.40 in TM3.<br />

Moreover, <strong>the</strong> H1R contains a polar N5.46 amino acid at position 5.46, which is involved in<br />

histamine binding (Leurs, Smit et al. 1994), linking this GPCR sub-domain directly to <strong>the</strong> agonist<br />

binding site.<br />

Influence <strong>of</strong> T3.37 and I3.40 in agonist-induced activation <strong>of</strong> <strong>the</strong> human histamine H1R<br />

In order to test <strong>the</strong> role <strong>of</strong> <strong>the</strong> amino acids at positions 3.37 and 3.40 we engineered mutants in<br />

which T3.37 <strong>of</strong> <strong>the</strong> histamine H1R was substituted by ei<strong>the</strong>r Ala (T3.37A), to remove <strong>the</strong> hydrogen<br />

bonding capability at this position, or Glu (T3.37E), to mimic <strong>the</strong> bovine rhodopsin sequence; and<br />

I3.40 was replaced by ei<strong>the</strong>r Ala (I3.40A) or Gly (I3.40G) removing <strong>the</strong> hydrophobic and bulky side<br />

chain and, thus, <strong>the</strong> van der Waals interaction with <strong>the</strong> carbonyl group at position 5.46. Figure 2E<br />

shows <strong>the</strong> environment <strong>of</strong> T3.37 in <strong>the</strong> β2-based molecular model <strong>of</strong> <strong>the</strong> histamine H1R (see<br />

Experimental Procedures). T3.37 hydrogen bonds N5.46, <strong>the</strong> binding partner for <strong>the</strong> imidazole ring<br />

<strong>of</strong> histamine (Leurs, Smit et al. 1994), in a similar manner as <strong>the</strong> T3.37•••S5.46 interaction in <strong>the</strong><br />

structure <strong>of</strong> <strong>the</strong> β2AR (Figure 1E). Wild type, T3.37A and T3.37E mutants H1Rs were well expressed<br />

in COS-7 cells (Bmax 14.8±4.6 (n=7), 8.5±2.1 (n=2) and 12.8±3.3 (n=2) pmol/mg protein,<br />

respectively) as determined by radioligand binding analysis (Table 2).<br />

93 | P a g e


Chapter 5<br />

The function <strong>of</strong> <strong>the</strong> WT H1R was evaluated in a NF-κB reporter-gene assay as previously reported<br />

(Jongejan, Bruysters et al. 2005; Bakker, Jongejan et al. 2008), and <strong>the</strong> activation <strong>of</strong> NF-κB was<br />

increased up to 10 fold when stimulated with histamine (pEC50=6.5±0.1) (Table 2 and Figure 2A).<br />

Removal <strong>of</strong> <strong>the</strong> interaction <strong>of</strong> T3.37 with N5.46, by mutating residue T3.37 to Ala resulted in a 4-<br />

fold decrease <strong>of</strong> <strong>the</strong> binding affinity <strong>of</strong> histamine (pKi <strong>of</strong> 5.3 for <strong>the</strong> WT H1R vs. 4.7 for <strong>the</strong> T3.37A<br />

H1R) (Table 2) as measured by [ 3H]mepyramine displacement. A lower constitutive signaling <strong>of</strong> <strong>the</strong><br />

T3.37A H1R mutant compared to that <strong>of</strong> WT H1R was observed (Figure 2A and 2B). However, <strong>the</strong><br />

basal signaling could still be increased by histamine albeit at higher concentrations (pEC50=4.5±0.2)<br />

(Table 2).<br />

Mutation <strong>of</strong> residue T3.37 to Glu (i.e. making this region similar to bovine rhodopsin) did not<br />

influence <strong>the</strong> binding affinity <strong>of</strong> histamine (pKi <strong>of</strong> 5.3 for <strong>the</strong> WT H1R versus 5.1 for <strong>the</strong> T3.37E H1R),<br />

possibly because E3.37 maintains both <strong>the</strong> interaction with N5.46 (Figure 2F), as in <strong>the</strong> WT H1R<br />

(Figure 2E), and with <strong>the</strong> carbonyl group at position 5.46, as in bovine rhodopsin (Figure 1C).<br />

Notably, <strong>the</strong> T3.37E mutant H1R hardly shows any constitutive activity (Figure 2A and 2B).<br />

Moreover, <strong>the</strong> T3.37E mutant H1R is activated very poorly by histamine (pEC50>3.5) (Table 2).<br />

Thus, we propose that <strong>the</strong> additional constraint between E3.37 and <strong>the</strong> backbone carbonyl at<br />

position 5.46, introduced in <strong>the</strong> T3.37E mutation, impedes <strong>the</strong> activation <strong>of</strong> <strong>the</strong> mutant H1R by<br />

locking a local structure that cannot be overcome by <strong>the</strong> agonist histamine. Our results also show<br />

that removal <strong>of</strong> <strong>the</strong> bulky hydrophobic I3.40 side chain by substitution to Ala or Gly is translated<br />

into significant effects on both <strong>the</strong> constitutive activity <strong>of</strong> <strong>the</strong> receptor and <strong>the</strong> process <strong>of</strong> histamine-<br />

induced receptor activation (Figure 2A and 2C). Importantly, <strong>the</strong>se effects are not due to a loss <strong>of</strong><br />

binding affinity for histamine. As shown in Table 2 both I3.40A and I3.40G mutant H1Rs are well<br />

expressed in COS-7 cells (Bmax 9.4±5.1 and 8.8±4.8 pmol/mg protein, respectively) and bind<br />

histamine with comparable affinity (pKi = 5.3±0.2 for I3.40A H1R and 5.0±0.2 for <strong>the</strong> I3.40G H1R)<br />

compared to WT H1R (pKi = 5.3±0.1).<br />

The fact that removal <strong>of</strong> <strong>the</strong> I3.40 side chain abolished both <strong>the</strong> constitutive and histamine-induced<br />

GPCR activity strongly suggests that this hydrophobic side chain plays a key role in <strong>the</strong> mechanism<br />

<strong>of</strong> activation. Based on <strong>the</strong>se results and on <strong>the</strong> recent insights in GPCR structure (Park, Scheerer et<br />

al. 2008; Scheerer, Park et al. 2008), we propose that <strong>the</strong> hydrophobic side chain at position 3.40,<br />

which is strongly packed against TM5 (Figure 1), acts as a pivot and <strong>the</strong> Pro5.50-induced<br />

unwinding acts as a hinge in <strong>the</strong> crucial reorientation <strong>of</strong> TM5 upon activation, which ultimately<br />

94 | P a g e


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

facilitates <strong>the</strong> interactions <strong>of</strong> Y5.58 and K5.66 with R3.50 and E6.30, respectively (Figure 1).<br />

Notably, mutations at position 3.40 in rhodopsin are related to retinitis pigmentosa, poor retinal<br />

binding, misfolding, reduced expression, decreased time in <strong>the</strong> MII state, and reduced transducin<br />

activation (Madabushi, Gross et al. 2004).<br />

Table 2 Pharmacological characteristics <strong>of</strong> hH1R wild type (WT) and H1R mutants transiently<br />

expressed in COS-7 cells.<br />

The potency (pEC50) <strong>of</strong> histamine was measured by NF-κB reporter gene assay. Affinities <strong>of</strong> [ 3 H]mepyramine<br />

(KD) and expression levels are determined by saturation radioligand binding assays. Binding affinity values <strong>of</strong><br />

H1R ligands are determined by [ 3 H]mepyramine displacement. Data were calculated as <strong>the</strong> mean ± S.E.M. (n)<br />

is number <strong>of</strong> experiments, each performed in triplicate.<br />

Receptor<br />

KD (nM)± S.E.M.<br />

(n)<br />

[ 3H]mepyramine Histamine<br />

Bmax(pmol/mg<br />

protein)<br />

± S.E.M. (n)<br />

pKi± S.E.M. (n) pEC50± S.E.M. (n)<br />

hH1R WT 1.5 ± 0.2 (7) 14.8 ± 4.6 (7) 5.3 ± 0.1 (4) 6.5 ± 0.1 (5)<br />

T3.37A 4.8 ± 0.3 (2) 8.5 ± 2.1 (2) 4.7 ± 0.1 (4) 4.5 ± 0.2 (3)<br />

T3.37E 6.2 ± 0.1 (2) 12.8 ± 3.3 (2) 5.1 ± 0.1 (4) > 3.2 (3)<br />

S3.39A 1.6 ± 0.3 (3) 8.9 ± 2.4 (3) 5.4 ± 0.1 (4) 4.9 ± 0.1 (3)<br />

S3.39C 1.5 ± 0.3 (3) 9.0 ± 3.2 (3) 5.3 ± 0.1 (4) 5.2 ± 0.1 (3)<br />

S3.39T 3.6 ± 0.3 (3) 8.2 ± 4.7 (3) 5.4 ± 0.1 (4) 4.9 ± 0.2 (3)<br />

I3.40A 10.1 ± 0.3 (3) 9.4 ± 5.1 (3) 5.3 ± 0.2 (4) > 3.4 (3)<br />

I3.40G 7.5 ± 0.2 (2) 8.8 ± 4.8 (2) 5.0 ± 0.2 (3) > 3.8 (2)<br />

S3.36T+<br />

I3.40A<br />

S3.36T+<br />

I3.40G<br />

7.8 ± 0.3 (4) 10.4 ± 5.3 (4) 5.4 ± 0.1 (3) > 3.7 (2)<br />

3.4 ± 0.2 (3) 13.4 ± 5.0 (3) 5.0 ± 0.1 (3) > 3.9 (2)<br />

To substantiate this hypo<strong>the</strong>sis we performed an additional set <strong>of</strong> experiments. We have recently<br />

shown that <strong>the</strong> S3.36T mutation in <strong>the</strong> histamine H1R induces a large increase in constitutive<br />

activity (Jongejan, Bruysters et al. 2005). We proposed that T3.36 induces <strong>the</strong> transition <strong>of</strong> W6.48<br />

toward TM5 (Jongejan, Bruysters et al. 2005; Pellissier, Sallander et al. 2009), considered to be <strong>the</strong><br />

initial stage <strong>of</strong> <strong>the</strong> activation process, as observed in <strong>the</strong> electron microscopy density map <strong>of</strong><br />

95 | P a g e


Chapter 5<br />

metarhodopsin I (Ruprecht, Mielke et al. 2004) and in solid-state NMR measurements <strong>of</strong><br />

metarhodopsin II (Crocker, Eilers et al. 2006). Thus, we combined <strong>the</strong> activating S3.36T point<br />

mutation with <strong>the</strong> inactivating I3.40A/G mutation to generate <strong>the</strong> S3.36T/I3.40A and<br />

S3.36T/I3.40G double mutant H1Rs, with <strong>the</strong> aim to assess <strong>the</strong>ir compensatory consequences on<br />

receptor function. The double mutant H1Rs are well expressed in COS-7 cells (Bmax 10.4±5.3 and<br />

13.4±0.3 pmol/mg protein, respectively) and have similar affinity for histamine compared to wild-<br />

type H1R (pKi 5.4±0.1 for <strong>the</strong> S3.36T/I3.40A H1R and 5.0±0.1 for <strong>the</strong> S3.36T/I3.40G H1R,<br />

respectively) (Table 2).<br />

Interestingly, both double mutants show lack <strong>of</strong> constitutive activity and cannot be activated by<br />

histamine, resembling <strong>the</strong> single I3.40A/G mutant H1R (Figure 2A and 2C). These data reinforce<br />

<strong>the</strong> importance <strong>of</strong> I3.40 in <strong>the</strong> process <strong>of</strong> receptor activation.<br />

The role <strong>of</strong> S3.36 and S3.39 in <strong>the</strong> process <strong>of</strong> histamine H1R activation<br />

S3.39, within <strong>the</strong> TAS(I/V) motif, is highly conserved in <strong>the</strong> biogenic amine (100% <strong>of</strong> <strong>the</strong><br />

sequences) and o<strong>the</strong>r (S:75%; T:6%) GPCR families, which suggests a conserved structural and/or<br />

functional role. S3.39, absent in bovine rhodopsin, is hydrogen bonding, in <strong>the</strong> g+ conformation, <strong>the</strong><br />

backbone carbonyl at position 3.35 in <strong>the</strong> previous turn <strong>of</strong> <strong>the</strong> helix in <strong>the</strong> structures <strong>of</strong> β1AR (at <strong>the</strong><br />

inter-atomic distance between heteroatoms <strong>of</strong> 3.0Å) and β2AR (2.7Å) adrenergic receptors and<br />

adenosine A2A receptor (3.1Å). S3.39 does not interact with <strong>the</strong> near and highly conserved D2.50 in<br />

TM2 (4.7Å, 4.1Å and 4.6Å in <strong>the</strong> β1AR, β2AR and A2A receptors, respectively) or near water<br />

molecules.<br />

Thus, to study <strong>the</strong> role <strong>of</strong> this amino acid on <strong>the</strong> receptor structure and function (see Table 1) we<br />

mutated S3.39 to alanine, threonine and cysteine. The S3.39A mutant H1R is well expressed in COS-<br />

7 cells (Bmax 8.9±2.4 pmol/mg protein (n=3)) and has a similar binding affinity for histamine (pKi<br />

5.4±0.1 (n=4)) compared to <strong>the</strong> wild-type H1R (Table 2). As shown in <strong>the</strong> NF-κB reporter gene<br />

assay (see Experimental Procedures), <strong>the</strong> S3.39A mutant H1R shows almost no constitutive activity<br />

(Figure 3A). However, <strong>the</strong> S3.39A mutant H1R could still be activated by histamine with pEC50 <strong>of</strong><br />

4.9±0.1 (n=3), but <strong>the</strong> maximal activity is reduced by a factor <strong>of</strong> 2 (Figure 3A and 3B). These<br />

results indicate that <strong>the</strong> hydrogen bond capability <strong>of</strong> S3.39 plays an important role in stabilizing <strong>the</strong><br />

active state <strong>of</strong> <strong>the</strong> histamine H1R. Fur<strong>the</strong>rmore, S3.39T/C mutations provide a way <strong>of</strong> identifying<br />

96 | P a g e


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

<strong>the</strong> conformation <strong>of</strong> S3.39 in <strong>the</strong> active state because <strong>of</strong> <strong>the</strong> different side chain rotamer<br />

distribution <strong>of</strong> serine from those <strong>of</strong> threonine or cysteine, when located in an α-helical structure<br />

(see Experimental Procedures, Table 1).<br />

Figure 1 The local opening <strong>of</strong> transmembrane helix 5 in bovine rhodopsin (PDB code 1GZM),<br />

opsin (3CAP) and <strong>the</strong> β2-adrenergic receptor (2RH1)<br />

(A, B) Evolution <strong>of</strong> bend (A) and helical twist (B) angles (°, see Experimental Procedures) along TM5 in <strong>the</strong><br />

crystal structures <strong>of</strong> <strong>the</strong> β2-adrenergic receptor (blue), bovine rhodopsin (black) and opsin (red). Residue<br />

number refers to <strong>the</strong> first residue in each turn; i.e. <strong>the</strong> peak <strong>of</strong> <strong>the</strong> helical distortion appears in <strong>the</strong> turn 5.45-<br />

5.48, labeled as 5.45 in <strong>the</strong> graphic. (C, D, E) Detailed view <strong>of</strong> <strong>the</strong> interface between TMs 3 (dark red) and 5<br />

(green) in bovine rhodopsin (C), opsin (D), and <strong>the</strong> �2-adrenergic receptor (E) in views parallel (top) and<br />

perpendicular (bottom) to <strong>the</strong> membrane.<br />

97 | P a g e


Chapter 5<br />

Fur<strong>the</strong>rmore, both S3.39T/C mutant H1Rs also lack constitutive activity and <strong>the</strong>ir maximal activites<br />

are also reduced by a factor <strong>of</strong> 2 relative to WT (Figure 3A and 3B). These results can be explained<br />

by an increase <strong>of</strong> <strong>the</strong> population <strong>of</strong> <strong>the</strong> g+ side chain conformation, which corresponds to <strong>the</strong><br />

inactive state (see above). These findings suggest that <strong>the</strong> side chain at position 3.39 changes from<br />

<strong>the</strong> inactive g+ to ei<strong>the</strong>r <strong>the</strong> g- or t active conformation upon activation. In Class A GPCRs position<br />

3.36 is highly variable; however, in amine receptors it is mostly C (57%) or S (32%), with <strong>the</strong><br />

notable exception <strong>of</strong> β-adrenergic receptors, which feature V. Remarkably, T is virtually absent<br />

(2%). The S3.36T mutation showed a large increase in constitutive activity (Figure 2A and 2D),<br />

suggesting that <strong>the</strong> side chain at this position plays a key role in stabilizing <strong>the</strong> active state <strong>of</strong> <strong>the</strong><br />

receptor in <strong>the</strong> g+ conformation, pointing towards TM7 (Jongejan, Bruysters et al. 2005).<br />

This conformational transition <strong>of</strong> S3.36 and <strong>the</strong> rotation <strong>of</strong> TM3 are only feasible if <strong>the</strong> side chain <strong>of</strong><br />

W6.48 modifies its conformation simultaneously, in a concerted manner, from <strong>the</strong> inactive g+<br />

(pointing towards TM7) to <strong>the</strong> active t conformation (pointing towards TM5) (Jongejan, Bruysters<br />

et al. 2005; Pellissier, Sallander et al. 2009). Figure 3C and 3D depict <strong>the</strong> conformational transition<br />

<strong>of</strong> <strong>the</strong>se side chains and <strong>the</strong> proposed interactions in <strong>the</strong> inactive and active states.<br />

The rotamer toggle switch <strong>of</strong> W6.48 relocates <strong>the</strong> indole ring <strong>of</strong> <strong>the</strong> side chain from pointing<br />

towards TM7, where it is engaged in a conserved, water-mediated, hydrogen bond interaction with<br />

D2.50 (Figure 3C) (Pardo, Deupi et al. 2007), to point towards <strong>the</strong> binding site crevice (Figure 3D)<br />

(Pardo, Deupi et al. 2007; Pellissier, Sallander et al. 2009).<br />

The rotamer toggle switch <strong>of</strong> S3.36 relocates its polar side chain from <strong>the</strong> binding pocket (Figure<br />

3C) (Jongejan, Bruysters et al. 2005) to point towards TM7 where it might fulfill a hydrogen bond<br />

interaction with <strong>the</strong> water molecules in <strong>the</strong> D2.50/W6.48 environment (Figure 3D). Finally, S3.39<br />

also switches from <strong>the</strong> inactive g+, hydrogen bonding <strong>the</strong> backbone carbonyl in <strong>the</strong> previous turn <strong>of</strong><br />

<strong>the</strong> helix (Figure 3C), to point towards TM7 where it stabilizes <strong>the</strong> active state <strong>of</strong> <strong>the</strong> receptor by<br />

interacting with <strong>the</strong> water molecules in <strong>the</strong> D2.50/W6.48 environment (Figure 3D).<br />

98 | P a g e<br />

DISCUSSION<br />

A number <strong>of</strong> studies have provided convincing evidence that GPCRs co-exist in different<br />

conformations (Kobilka and Deupi 2007). GPCRs are maintained within <strong>the</strong> ensemble <strong>of</strong> inactive


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

conformations through non-covalent interactions between side chains, mostly located in <strong>the</strong> TM<br />

segments (Smit, Vischer et al. 2007; Weis and Kobilka 2008; Nygaard, Frimurer et al. 2009).<br />

Disruption <strong>of</strong> <strong>the</strong>se stabilizing interactions by Ala-scanning mutagenesis leads to gain-<strong>of</strong>-function<br />

mutations. The crystal structures <strong>of</strong> <strong>the</strong> inactive state <strong>of</strong> bovine (Li, Edwards et al. 2004) and squid<br />

(Murakami and Kouyama 2008) rhodopsin, adenosine A2A receptor (Jaakola, Griffith et al. 2008),<br />

and <strong>the</strong> β1AR (Warne, Serrano-Vega et al. 2008) and β2AR (Cherezov, Rosenbaum et al. 2007;<br />

Rasmussen, Choi et al. 2007) facilitate <strong>the</strong> structural interpretation <strong>of</strong> <strong>the</strong>se experimental results.<br />

However, understanding <strong>the</strong> mechanisms that shift <strong>the</strong> equilibrium <strong>of</strong> <strong>the</strong> ensemble to <strong>the</strong> active<br />

conformations is more challenging for two reasons. First, alanine-scanning mutagenesis <strong>of</strong> <strong>the</strong> side<br />

chains involved in <strong>the</strong> stabilization <strong>of</strong> active conformations leads to loss-<strong>of</strong>-function effects (Table<br />

1), which are more difficult to measure with precision. Second, high-resolution structural<br />

information <strong>of</strong> <strong>the</strong> active state is very scarce, currently limited to <strong>the</strong> cytoplasmic side <strong>of</strong> <strong>the</strong> recent<br />

crystal structure <strong>of</strong> <strong>the</strong> ligand-free opsin (Park, Scheerer et al. 2008), which hampers <strong>the</strong> structural<br />

interpretation <strong>of</strong> <strong>the</strong> experimental results. The fact that <strong>the</strong> histamine H1R shows considerable<br />

agonist-independent, constitutive signaling, which can be strongly reduced by <strong>the</strong> inverse agonist<br />

mepyramine (Bakker, Schoonus et al. 2001), provides an unique opportunity to approach <strong>the</strong><br />

mechanism <strong>of</strong> receptor activation. We have designed point mutations that modulate <strong>the</strong> strenght <strong>of</strong><br />

intramolecular interactions and <strong>the</strong> side chain rotamer distribution <strong>of</strong> specific residues, and<br />

measured <strong>the</strong>ir effect on constitutive agonist-independent activity (see Table 1).<br />

Specifically, we have focused on <strong>the</strong> highly conserved TAS(I/V) motif in TM3 <strong>of</strong> amine receptors. An<br />

evolutionary trace analysis has led to <strong>the</strong> identification <strong>of</strong> a putative common signal transduction<br />

process for Class A GPCRs, where key residues are grouped into a network <strong>of</strong> interactions extending<br />

from <strong>the</strong> binding pocket to <strong>the</strong> cytoplasmic region (Madabushi, Gross et al. 2004). This analysis<br />

detected L3.40 <strong>of</strong> rhodopsin as a structurally and/or functionally important residue.<br />

Our results support a role <strong>of</strong> this amino acid in <strong>the</strong> process <strong>of</strong> histamine H1R activation. Mutation <strong>of</strong><br />

I3.40 to ei<strong>the</strong>r Ala or Gly, removing <strong>the</strong> bulky side chain, abolished <strong>the</strong> constitutive activity <strong>of</strong> <strong>the</strong><br />

receptor, <strong>the</strong> effect <strong>of</strong> constitutive-activity increasing mutations, and <strong>the</strong> histamine-induced<br />

receptor activation. In <strong>the</strong> inactive state (Figure 1C and 1E) this hydrophobic side chain at position<br />

3.40 (green surface) is located between <strong>the</strong> pyrrolidine ring <strong>of</strong> P5.50 and <strong>the</strong> carbonyl oxygen at<br />

position 5.46, inducing or stabilizing <strong>the</strong> local opening <strong>of</strong> TM5 (Figure 1B).<br />

99 | P a g e


Chapter 5<br />

Figure 2 Influence <strong>of</strong> T3.37 and I3.40 in <strong>the</strong> agonist-induced activation <strong>of</strong> <strong>the</strong> histamine H1<br />

receptor. NF-κB activation modulated by histamine in COS-7 cells transiently transfected with<br />

wild type (WT), T3.37, I3.40, S3.36, double H1R mutants and empty vector (pcDEF3)<br />

(A) Basal activity (black bars) and after stimulation with 10-4 M histamine (grey bars). Results are normalized to<br />

<strong>the</strong> basal activity <strong>of</strong> WT which is put as 100%. (B) Representative dose response curves <strong>of</strong> histamine for WT (●),<br />

T3.37A (○), T3.37E (�), and pcDEF3 (□). (C) Representative dose response curves <strong>of</strong> histamine for WT (●), I3.40A<br />

(○), I3.40G (�), and pcDEF3 (□). (D) Representative dose response curves <strong>of</strong> histamine for WT (●), S3.36T (○),<br />

S3.36T+I3.40G (�), S3.36T+I3.40A (♦) and pcDEF3 (□). (E, F) Detailed view <strong>of</strong> <strong>the</strong> interface between TMs 3 and 5<br />

in <strong>the</strong> molecular model <strong>of</strong> WT (E) and mutant T3.37E (F) histamine H1R.<br />

100 | P a g e


A<br />

B<br />

C<br />

D<br />

Figure 3 Influence <strong>of</strong><br />

S3.39 in <strong>the</strong> agonistinduced<br />

activation <strong>of</strong><br />

<strong>the</strong> histamine H1<br />

receptor<br />

(A) Basal activity (black<br />

bars) and after stimulation<br />

with 10 -4 M histamine (grey<br />

bars). Results are<br />

normalized to <strong>the</strong> basal<br />

activity <strong>of</strong> WT which is put<br />

as 100%. (B)<br />

Representative dose<br />

response curves <strong>of</strong><br />

histamine for WT (•),<br />

S3.39T (ο), S3.39A (�),<br />

S3.39C ( ) hH1R mutants<br />

and pcDEF3 (♦). (C, D)<br />

Proposed hydrogen-bond<br />

network involving <strong>the</strong> side<br />

chains <strong>of</strong> D2.50, S3.36,<br />

S3.39 and W6.48, and <strong>the</strong><br />

water molecules in <strong>the</strong><br />

D2.50/W6.48 environment<br />

in <strong>the</strong> inactive (C) and<br />

active (D) models <strong>of</strong> <strong>the</strong><br />

histamine H1R.


Chapter 5<br />

However, <strong>the</strong> molecular mechanism by which this residue facilitates receptor activation is not<br />

straightforward. The structure <strong>of</strong> opsin (Park, Scheerer et al. 2008) and previous substituted-<br />

cysteine accessibility studies in <strong>the</strong> β2AR (Ge<strong>the</strong>r, Lin et al. 1997) and µ- opioid receptor(Shi,<br />

Liapakis et al. 2002) have shown a clockwise rotation (viewed from <strong>the</strong> intracellular side) <strong>of</strong> TM3<br />

during <strong>the</strong> process <strong>of</strong> receptor activation. Rotation <strong>of</strong> TM3 moves <strong>the</strong> bulky side chain at position<br />

3.40, located between P5.50 and <strong>the</strong> 5.46 carbonyl, away from <strong>the</strong>se moieties (Figure 1D), resulting<br />

in a localized decrease <strong>of</strong> <strong>the</strong> bend around P5.50 (Figure 1A).<br />

Fluorescence spectroscopy experiments used to monitor agonist-induced conformational changes<br />

in <strong>the</strong> β2AR provide evidence <strong>of</strong> a sequential binding model through discrete intermediate<br />

conformational states (Yao, Parnot et al. 2006). In an initial step agonist binding induces relocation<br />

<strong>of</strong> <strong>the</strong> extracellular side <strong>of</strong> TM5 to facilitate binding <strong>of</strong> <strong>the</strong> catechol hydroxyls <strong>of</strong> <strong>the</strong> β2AR ligand<br />

(Kobilka 2004). We propose that this movement is transduced as a set <strong>of</strong> structural rearrangements<br />

toward <strong>the</strong> intracellular side <strong>of</strong> TM5 with <strong>the</strong> assistance <strong>of</strong> I3.40 acting as a pivot (Figure 2C),<br />

ultimately enabling <strong>the</strong> interactions <strong>of</strong> Y5.58 with R3.50 and K5.66 with E6.30 in <strong>the</strong> cytoplasmic<br />

side <strong>of</strong> <strong>the</strong> receptor.<br />

In support <strong>of</strong> this hypo<strong>the</strong>sis, <strong>the</strong> T3.37E mutation provides an additional constraint between TMs<br />

3 and 5 (Figure 2F), <strong>the</strong>reby preventing <strong>the</strong> transmission <strong>of</strong> <strong>the</strong> activation signal through TM5<br />

(Figure 2B). In addition to this activation pathway through TM5, agonist binding triggers<br />

additional routes <strong>of</strong> activation, for instance by activating <strong>the</strong> rotamer toggle switch <strong>of</strong> tryptophan at<br />

position 6.48 (W6.48) from <strong>the</strong> inactive g+ (pointing towards TM7) to <strong>the</strong> active t (towards <strong>the</strong><br />

binding pocket) conformation (Shi, Liapakis et al. 2002; Ruprecht, Mielke et al. 2004; Crocker, Eilers<br />

et al. 2006; Ahuja, Crocker et al. 2009) (Figure 3D and 3E), through <strong>the</strong> formation <strong>of</strong> specific<br />

hydrogen bonds (Pardo, Deupi et al. 2007; Pellissier, Sallander et al. 2009).<br />

Notably, <strong>the</strong> rotamer toggle switch <strong>of</strong> W6.48 occurs in a concerted manner with <strong>the</strong> side chain at<br />

position 3.36 from <strong>the</strong> inactive t (towards <strong>the</strong> binding pocket) to <strong>the</strong> active g+ (towards TM7)<br />

conformation (Figure 3D and 3E) (Jongejan, Bruysters et al. 2005; Pellissier, Sallander et al. 2009).<br />

In this manuscript we have first shown that S3.39 also switches from <strong>the</strong> inactive g+ (hydrogen<br />

bonding <strong>the</strong> backbone carbonyl in <strong>the</strong> previous turn <strong>of</strong> <strong>the</strong> helix) to <strong>the</strong> active g- or t conformer,<br />

pointing towards TM7 and stabilizing <strong>the</strong> active state <strong>of</strong> <strong>the</strong> receptor probably by interacting with<br />

<strong>the</strong> water molecules in <strong>the</strong> D2.50/W6.48 environment (Figure 3E).<br />

102 | P a g e


A structural insight into <strong>the</strong> reorientation <strong>of</strong> TM3 and TM5 during H1R activation<br />

In <strong>the</strong> H1R, histamine binding induces and/or stabilizes <strong>the</strong> conformational switches <strong>of</strong> S3.36, S3.39,<br />

and W6.48 (toggle switch), triggering <strong>the</strong> conformational transition <strong>of</strong> N7.49 (<strong>the</strong> NPxxY motif <strong>of</strong><br />

TM7) towards D2.50 (Urizar, Claeysen et al. 2005) (Figure 4), and ultimately lead to <strong>the</strong> disruption<br />

<strong>of</strong> <strong>the</strong> ionic interaction between R3.50 and <strong>the</strong> adjacent D/E3.49 (<strong>the</strong> DRY motif <strong>of</strong> TM3) (Scheer,<br />

Fanelli et al. 1996), and an additional D/E6.30 near <strong>the</strong> cytoplasmic end <strong>of</strong> TM6 (Govaerts, Lefort et<br />

al. 2001) that is not present in several GPCR families (Springael, de Poorter et al. 2007). Disruption<br />

<strong>of</strong> <strong>the</strong>se ionic interactions permit R3.50 to adopt an extended conformation (Bakker, Jongejan et al.<br />

2008; Park, Scheerer et al. 2008), pointing toward <strong>the</strong> protein core, to interact with <strong>the</strong> highly<br />

conserved Y5.58 and Y7.53 in TMs 5 and 7 (Park, Scheerer et al. 2008).<br />

A recent computer model analysis <strong>of</strong> <strong>the</strong> activation mechanism <strong>of</strong> <strong>the</strong> guinea pig H1R found<br />

somewhat similar findings. The model proposes that during <strong>the</strong> first stage <strong>of</strong> receptor activation,<br />

straightening <strong>of</strong> TM VI, causes a rearrangment bulky aminoacids in TM VI. This <strong>the</strong>n causes <strong>the</strong><br />

interaction between <strong>of</strong> S3.39 that interacts with N7.45 to be lost. After activation S3.39 moves to<br />

TM2 to form a water mediated hydrogen bond with D2.50. Unlike <strong>the</strong> human version, in <strong>the</strong> guinea<br />

pig H1R <strong>the</strong>re is only a very limited role for S3.36 in <strong>the</strong> activation process (Strasser, Wittmann et al.<br />

2007).<br />

Thus, <strong>the</strong>re is an activation pathway leading from agonist binding to <strong>the</strong> cytoplasmic side <strong>of</strong> <strong>the</strong><br />

receptor through TMs 3, 6 and 7 (Figure 4, left arrows). Fluorescence spectroscopy in <strong>the</strong> β2AR has<br />

shown that partial agonists are as effective as full agonists in disrupting <strong>the</strong> ionic interaction <strong>of</strong><br />

R3.50, <strong>the</strong>refore, additional pathways <strong>of</strong> signal propagation must exist (Yao, Parnot et al. 2006). We<br />

propose that one <strong>of</strong> such pathways consists in <strong>the</strong> agonist-induced relocation <strong>of</strong> <strong>the</strong> extracellular<br />

side <strong>of</strong> TM5 towards <strong>the</strong> bundle, which is transmitted towards <strong>the</strong> cytoplasmic region through <strong>the</strong><br />

fulcrum/lever mechanism described above, in which P5.50 acts as a hinge and <strong>the</strong> highly conserved<br />

hydrophobic I3.40 side chain acts as a pivot (Figure 4, right arrow). Triggering specific pathways<br />

along <strong>the</strong> activation process might be an explanation for <strong>the</strong> observed collateral efficacy (Kenakin<br />

2007) <strong>of</strong> GPCR agonists. A thorough understanding how a ligand-encoded extracellular signal is<br />

propagated from its binding site into intracellular micro-domains will have a big impact on our<br />

understanding <strong>of</strong> <strong>the</strong> physiology and pathology <strong>of</strong> this important family <strong>of</strong> proteins, and in <strong>the</strong><br />

<strong>the</strong>rapeutic use <strong>of</strong> agonists and GPCR modulators in <strong>the</strong> clinic.<br />

103 | P a g e


Chapter 5<br />

Figure 4 Pathways <strong>of</strong> activation in <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs<br />

Agonist binding at <strong>the</strong> extracellular domain <strong>of</strong> <strong>the</strong> receptor leads to an activation pathway through TMs 3, 6<br />

and 7 (left arrows); or through relocation <strong>of</strong> TM5 assisted by <strong>the</strong> conserved hydrophobic side chain at<br />

position 3.40, depicted as a green sphere (right arrow).<br />

104 | P a g e


CHAPTER<br />

The role <strong>of</strong> isoleucine at position 6.45 in<br />

<strong>the</strong> H1R<br />

Kamonchanok Sansuk 1, Remko Bakker 1, 3, Anne Watts 1, Leonardo<br />

Pardo 2, Rob Leurs 1<br />

1Leiden/<strong>Amsterdam</strong> Center for Drug Research (LACDR), <strong>Vrije</strong> <strong>Universiteit</strong> <strong>Amsterdam</strong>,<br />

Department <strong>of</strong> Medicinal Chemistry, De Boelelaan 1083, 1081HV, <strong>Amsterdam</strong>, <strong>the</strong> Ne<strong>the</strong>rlands<br />

2Laboratori de Medicina Computacional, Unitat de Bioestadistica, Facultat de Medicina,<br />

Universitat Autonoma de Barcelona, 08193 Bellaterra, Spain<br />

3Current address: Boehringer Ingelheim Pharma GmbH&Co. KG, Biberach, Germany<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Chapter 6<br />

106 | P a g e<br />

Introduction<br />

Several mediators are involved in <strong>the</strong> patho-physiology <strong>of</strong> allergic diseases. However, histamine,<br />

which is syn<strong>the</strong>sized and released by human basophils, mast cells, lymphocytes, neurons and<br />

gastric enterochromaffin-like cells, remains <strong>the</strong> principal molecule that plays a fundamental role in<br />

<strong>the</strong> genesis <strong>of</strong> <strong>the</strong>se diseases, particularly rhinitis and urticaria. To date four subtypes <strong>of</strong> histamine<br />

receptors have been described (H1, H2, H3 and H4 receptors), which all belong to <strong>the</strong> super family <strong>of</strong><br />

G-protein coupled receptors (GPCRs). Histamine is thought to exert its effects in allergic disease<br />

primarily through <strong>the</strong> interaction with <strong>the</strong> histamine H1 receptor (H1R), which is present in a<br />

variety <strong>of</strong> cells (Leurs, Church et al. 2002; Akdis and Simons 2006). Consequently, <strong>the</strong> H1R is an<br />

important drug target, as during <strong>the</strong> past 20 years H1-receptor antagonists have become one <strong>of</strong> <strong>the</strong><br />

most frequently prescribed medications in Western countries to relieve <strong>the</strong> symptoms <strong>of</strong> allergic<br />

reactions (Pierce, Premont et al. 2002; Lundstrom 2005). The H1R exists in equilibrium between its<br />

inactive and active state. Histamine and o<strong>the</strong>r H1R agonists stabilize <strong>the</strong> active conformation and<br />

<strong>the</strong>reby promote H1R-mediated signal transduction. Even in <strong>the</strong> absence <strong>of</strong> H1R agonists <strong>the</strong> H1R<br />

can also be spontaneously active (Bakker, Schoonus et al. 2001). The constitutive activity <strong>of</strong> <strong>the</strong> H1R<br />

can be inhibited by H1R inverse agonists like mepyramine that stabilize <strong>the</strong> inactive conformation<br />

<strong>of</strong> <strong>the</strong> receptor and <strong>the</strong>reby block H1R-mediated signal transduction (Leurs, Church et al. 2002).<br />

Constitutive activity may also be induced by mutations <strong>of</strong> amino acid residues in <strong>the</strong> receptor.<br />

These constitutively active mutants have been successfully used for several GPCRs to gain insight in<br />

<strong>the</strong> mechanism <strong>of</strong> receptor activation (Ballesteros, Jensen et al. 2001; Govaerts, Lefort et al. 2001;<br />

Jongejan, Bruysters et al. 2005).<br />

In order to create constitutively active mutant H1R, we previously employed random mutagenesis<br />

in transmembrane domain (TM) regions known to be important for <strong>the</strong> receptor activation (Bakker,<br />

Jongejan et al. 2008). These regions included <strong>the</strong> highly conserved DRY motif in TM3, <strong>the</strong> top <strong>of</strong><br />

TM5 and <strong>of</strong> TM6. The subsequently generated mutants were initially screened using <strong>the</strong> receptor<br />

selection and amplification technology (R-SAT) and three residues in TM6: I420 (I6.40), A423<br />

(A6.43) and I425 (I6.45) were identified as sites prone to yield highly constitutively active mutants<br />

(Bakker, Jongejan et al. 2008). Fur<strong>the</strong>r characterization <strong>of</strong> <strong>the</strong>se constitutively active mutants in a<br />

nuclear factor-κ B (NF-κB) reporter gene assay led to <strong>the</strong> finding that <strong>the</strong> I6.45R H1R mutant<br />

exhibits an exceptional pharmacological pr<strong>of</strong>ile. Mutation <strong>of</strong> H1R at position 6.45 from isoleucine to<br />

arginine resulted in a unique receptor that no longer gives a NF-κB response when stimulated with


The role <strong>of</strong> isoleucine at position 6.45 in <strong>the</strong> H1R<br />

<strong>the</strong> endogenous ligand histamine. In this study, we <strong>the</strong>refore pharmacologically characterized <strong>the</strong><br />

I6.45R H1R mutant in detail in order to gain insights in <strong>the</strong> mechanism <strong>of</strong> <strong>the</strong> H1R activation.<br />

Materials<br />

Experimental procedures<br />

Gifts <strong>of</strong> mianserin hydrochloride (Schering-Plough, <strong>the</strong> Ne<strong>the</strong>rlands), pcDEF3 (Dr. J. Langer<br />

(Goldman, Cutrone et al. 1996)) and <strong>of</strong> <strong>the</strong> cDNA encoding <strong>the</strong> human histamine H1R (Dr. H. Fukui<br />

(Ohta, Hayashi et al. 1994)) are greatly acknowledged). All H1R mutants are available in <strong>the</strong> lab<br />

(Bakker, Jongejan et al. 2008). pNF-κB-Luc was obtained from Stratagene (La Jolla, USA), ATP<br />

disodium salt, bovine serum albumin, chloroquine diphosphate, DEAE-dextran (chloride form),<br />

histamine dihydrochloride, mepyramine (pyrilamine maleate) and polyethylenemine were<br />

purchased from Sigma Chemical Company (USA). D-luciferin was obtained from Duchefa Biochemie<br />

BV (Haarlem, The Ne<strong>the</strong>rlands), glycerol from Riedel-de-Haen (Germany), and Triton X-100 from<br />

Fluka (Switzerland). Cell culture media, penicillin and streptomycin were obtained from Invitrogen<br />

Corporation (Paisley, UK). Fetal bovine serum was obtained from Intergro B.V. (Dieren, The<br />

Ne<strong>the</strong>rlands). Cell culture plastics were obtained from Corning Costar (NY, USA). [ 3H]mepyramine<br />

(30 Ci/mmol) was purchased-from ICN Biomedicals BV (Zoetermeer, The Ne<strong>the</strong>rlands)<br />

Cell culture and transfection<br />

COS-7 African green monkey kidney cells were maintained at 37 °C in a humidified 5% CO2/95% air<br />

atmosphere in Dulbecco’s Modified Eagle’s Medium (DMEM) containing 50 IU/ml penicillin, 50<br />

µg/ml streptomycin and 5% (v/v) fetal bovine serum. COS-7 cells were transiently transfected<br />

using <strong>the</strong> DEAE-dextran method as previously described (Bakker, Schoonus et al. 2001).<br />

Reporter-gene assay<br />

Reporter-gene assays were performed essentially as described previously (Bakker, Schoonus et al.<br />

2001). In brief cells transiently co-transfected with pNF-κB-Luc (125 µg/1.10 7 cells) and pcDEF3<br />

encoding WT or mutant H1Rs (25 µg/1.10 7 cells) were seeded in 96 well white plates (Costar) in<br />

serum free culture medium and incubated with drugs. After 48 hrs, cells were assayed for<br />

luminescence by aspiration <strong>of</strong> <strong>the</strong> medium and <strong>the</strong> addition <strong>of</strong> 25 µl/well luciferase assay reagent<br />

(0.83 mM ATP. 0.83 mM d-luciferin, 18.7 mM MgCl2, 0.78 µM Na2H2P2O7, 38.9 mM Tris (pH 7.8).<br />

107 | P a g e


Chapter 6<br />

0.39% (v/v) glycerol, 0.03% (v/v) Triton X-100 and 2.6 µM dithiothreitol). After 30 min,<br />

luminescence was measured for 3 sec/well in a Victor plate reader (Wallac).<br />

Histamine H1R binding studies<br />

Histamine H1R binding studies were performed essentially as described previously (Bakker,<br />

Schoonus et al. 2001). In brief, transfected COS-7 cells used for radioligand binding studies were<br />

harvested after 48hrs and homogenized in ice-cold 50 mM Na2/K-phosphate buffer (pH 7.4)<br />

(binding buffer). The COS-7 cell homogenates were incubated for 30 min at 30 °C in binding buffer<br />

in 200 µl with 3 nM [ 3H]mepyramine. Non-specific binding was determined in <strong>the</strong> presence <strong>of</strong> 1 µM<br />

mianserin. The incubations were stopped by rapid dilution with 3 ml ice-cold binding buffer. The<br />

bound radioactivity was separated by filtration through Whatman GF/C filters that had been<br />

treated with 0.3% polyethyleneimine. Filters were washed twice with 3 ml binding buffer and<br />

radioactivity retained on <strong>the</strong> filters was measured by liquid scintillation counting.<br />

Analytical methods<br />

Protein concentrations were determined according to Bradford (Bradford 1976), using BSA as a<br />

standard. Binding and functional data were evaluated by a non-linear, least squares curve-fitting<br />

procedure using GraphPad Prism 4 ® (GraphPad S<strong>of</strong>tware, Inc., San Diego, CA).<br />

The role <strong>of</strong> I6.45 in <strong>the</strong> H1R binding pocket<br />

108 | P a g e<br />

Result and discussion<br />

Using random mutagenesis methodology, we previously identified several highly Constitutively<br />

Active Mutants (CAMs) <strong>of</strong> H1R upon mutation at position 6.40 and 6.45 in TM6 (Bakker et.al. 2008).<br />

The preliminary studies <strong>of</strong> <strong>the</strong> H1R mutants at <strong>the</strong> position 6.45 showed <strong>the</strong> unique<br />

pharmacological pr<strong>of</strong>ile <strong>of</strong> <strong>the</strong> I6.45R H1R mutant. The classical radioligand for H1R,<br />

[ 3H]mepyramine (see chemical structure at Figure 1) does not bind to <strong>the</strong> I6.45R H1R mutant.<br />

However, <strong>the</strong> tricyclic radioligand [ 3H]doxepin (see chemical structure at Figure 1), could bind <strong>the</strong><br />

<strong>the</strong> I6.45R H1R mutant with an affinity <strong>of</strong> ~ 8.0 nM.<br />

In this study, we continued to study <strong>the</strong> pharmacology <strong>of</strong> <strong>the</strong> I6.45R H1R mutant in more detail. As<br />

shown in Table 1, <strong>the</strong> wild type H1R was well expressed in COS-7 cells (Bmax 14.8±4.6 pmol/mg<br />

protein, n=7, p and KD 1.5±0.2 nM, n=7) as measured by radioligand [ 3H]mepyramine binding. In


The role <strong>of</strong> isoleucine at position 6.45 in <strong>the</strong> H1R<br />

contrary to <strong>the</strong> wild type H1R, <strong>the</strong> I6.45R H1R mutant lost its binding affinity towards <strong>the</strong><br />

radioligand [ 3H]mepyramine in a good agreement with <strong>the</strong> previous studies [Bakker et al. 2008].<br />

Mepyramine Doxepin Ketotifen<br />

Figure 1 Chemical structures <strong>of</strong> H1R inverse agonists: mepyramine, doxepin and ketotifen<br />

Table 1<br />

Pharmacological values <strong>of</strong> <strong>the</strong> wild type H1R and I6.45 H1R mutants<br />

Constructs<br />

[ 3H]mepyramine [ 3H]doxepin<br />

KD Bmax KD Bmax<br />

nM±S.E.M. (N) pmol/mg protein<br />

±S.E.M. (N)<br />

nM±S.E.M. (N)<br />

pmol/mg protein<br />

±S.E.M. (N)<br />

WT 1.5±0.2(7) 14.8±4.6(7) 1.5±0.2(3) 9.3±1.2(3)<br />

I 6.45R c.n.d. c.n.d. 10.4±0.1(3) 5.1±1.1(3)<br />

The values obtained from radioligand binding studies <strong>of</strong> <strong>the</strong> wild type H1R and I6.45R H1R mutant expressed<br />

in COS-7 cells using ei<strong>the</strong>r [ 3 H]mepyramine or [ 3 H]doxepin as a radiotracer. Data shown are mean values ±<br />

S.E.M. (n, number <strong>of</strong> experiments). c.n.d. = <strong>the</strong> value cannot be determined<br />

The I6.45 H1R mutant was, however, able to bind to <strong>the</strong> radioligand [ 3H]doxepin with Bmax and KD<br />

values <strong>of</strong> 5.1±1.1 pmol/mg protein (n=3) and 10.4±0.1 nM (n=3), respectively. Using [ 3H]doxepin as<br />

radiotracer, we observed that histamine binds with a high binding affinity for <strong>the</strong> I6.45R H1R<br />

mutant (pKi 6.9±0.2, n=4), compared to <strong>the</strong> wild type H1R (pKi 5.1±0.1, n=2), whereas mepyramine<br />

has a drastically decreased affinity for <strong>the</strong> I6.45R H1R mutant (pKi 5.2±0.2, N=3), compared to <strong>the</strong><br />

wild type H1R (pKi 9.2±0.1, N=2). Ketotifen (Figure 1), which is a H1R inverse agonist that contains a<br />

tricyclic moiety similar to doxepin, binds to both <strong>the</strong> wild type H1R and <strong>the</strong> I6.45R H1R mutant in a<br />

similar manner (Table 2 and Figure 2).<br />

109 | P a g e


Chapter 6<br />

Table 2<br />

pKi values <strong>of</strong> <strong>the</strong> wild type H1R and I6.45 H1R mutants<br />

Constructs<br />

110 | P a g e<br />

pKi values<br />

Histamine Mepyramine Doxepin Ketotifen<br />

WT 5.1±0.1 (2) 9.2±0.1 (2) 8.7±0.3(4) 9.1±0.1 (3)<br />

I 6.45R 6.9±0.2 (4) 5.2±0.2 (3) 8.2±0.4(5) 9.2±0.2 (4)<br />

The values obtained from radioligand binding studies <strong>of</strong> <strong>the</strong> wild type H1R and I6.45R H1R mutant expressed<br />

in COS-7 cells using [ 3 H]doxepin as a radiotracer. Data shown are mean values ± S.E.M. (n, number <strong>of</strong><br />

experiments). c.n.d. = <strong>the</strong> value cannot be determined<br />

Figure 2 [ 3H]doxepin binding analysis <strong>of</strong> <strong>the</strong> H1R wild type or I6.45R H1R mutant<br />

The radioligand binding assays were performed using COS-7 cells membranes transiently transfected with<br />

<strong>the</strong> H1R wild type (A) or I6.45R H1R mutant (B) in <strong>the</strong> presence <strong>of</strong> different concentrations <strong>of</strong> histamine (•),<br />

mepyramine (ο), doxepin (�) and ketotifen ( ). Data shown are from a representative experiments (N≥ 2)<br />

performed in triplicate.<br />

Although <strong>the</strong> residue I6.45 in TM6 <strong>of</strong> <strong>the</strong> H1R has previously not been implicated in <strong>the</strong> H1R ligand-<br />

binding site, our present results suggest that it is possible that <strong>the</strong> mutation at this position may<br />

change <strong>the</strong> structure <strong>of</strong> TM6 in such a way that it affect <strong>the</strong> binding <strong>of</strong> ligands. In addition to this,<br />

<strong>the</strong> ligand binding studies using [ 3H]mepyramine and [ 3H]doxepin have revealed that <strong>the</strong> I6.45R<br />

H1R mutant has low affinity for mepyramine, but still displays a high affinity for doxepin. The


The role <strong>of</strong> isoleucine at position 6.45 in <strong>the</strong> H1R<br />

disruption <strong>of</strong> <strong>the</strong> orientation <strong>of</strong> TM6 through <strong>the</strong> I6.45R H1R mutation might disturb an important<br />

interaction between mepyramine and F6.52, which does not affect <strong>the</strong> binding <strong>of</strong> doxepin. This<br />

indicates that <strong>the</strong> tricyclic compounds possibly have a different binding mode compared to <strong>the</strong><br />

biaryl class <strong>of</strong> H1R inverse agonists such as mepyramine. This is in good agreement with previous<br />

studies, where <strong>the</strong> binding <strong>of</strong> doxepin and related compounds was shown to be independent from<br />

an interaction with TM 6 and showed a crucial role for only D3.32 (Nonaka, Otaki et al. 1998).<br />

Figure 3 Basal activity <strong>of</strong> <strong>the</strong> Basal H1R activation and maximal histamine-induced NF-κB<br />

activation<br />

The basal activity (white bar) and maximal histamine-induced NF-κB activation (black bars) <strong>of</strong> <strong>the</strong> wild type<br />

H1R (WT), <strong>the</strong> I6.45R and I6.45N H1R mutant and <strong>the</strong> pcDEF3 vector (MOCK) in COS-7 cells. The basal<br />

activity <strong>of</strong> WT H1R is put at 100%. Values are mean± S.E.M. <strong>of</strong> at least 3 independent experiments.<br />

The role <strong>of</strong> I6.45 in <strong>the</strong> H1R activation<br />

The function <strong>of</strong> <strong>the</strong> wild type H1R was evaluated by <strong>the</strong> NF-κB reporter-gene assay as previously<br />

reported (Bakker, Jongejan et al. 2008). The wild type H1R is constitutively active and NF-κB<br />

activation is increased when stimulating <strong>the</strong> wild type H1R receptor with histamine (Figure 3).<br />

Mutation <strong>of</strong> <strong>the</strong> isoleucine at position 6.45 to arginine results in a H1R mutant (I6.45R) that lost its<br />

constitutive activity. Interestingly, despite its high affinity for histamine, <strong>the</strong> I6.45R H1R mutant<br />

could not be activated by histamine.<br />

111 | P a g e


Chapter 6<br />

Figure 4 Modulation <strong>of</strong> NF-κB and CRE activity <strong>the</strong> wild type H1R or <strong>the</strong> I6.45R H1R mutant<br />

The reporter gene assays were performed using COS-7 cells expressing <strong>the</strong> wild type H1R or <strong>the</strong> I6.45R H1R<br />

mutant. The transfected cells were induction by H1R ligands such as histamine (�), mepyramine ( ), doxepin (•)<br />

and ketotifen (ο) after 48 hrs <strong>of</strong> transfection. Basal activity <strong>of</strong> each construct is put at 100%. Data shown are<br />

from a representative experiments (N≥ 2) performed in triplicate.<br />

Even more surprising, H1R inverse agonists that contain a tricyclic moiety such as doxepin and<br />

ketotifen, were able to stimulate this I6.45R H1R mutant as agonists with high potencies <strong>of</strong> 7.2±0.1<br />

(N=7) and 8.8±0.3 (N=7), respectively (Table 3 and Figure 4). Similar effects <strong>of</strong> histamine and<br />

<strong>the</strong>se tricyclic H1R ligands with <strong>the</strong> wild type H1R and I6.45R H1R mutant were also observed in a<br />

CRE reporter gene assay (Table 3 and Figure 4). In CRE reporter gene assay, doxepin and ketotifen<br />

were able to stimulate <strong>the</strong> I6.45R H1R mutant with potencies <strong>of</strong> 7.7±0.3 (N=3) and 8.9±0.2 (N=4),<br />

respectively. Since reporter gene assays to trace activation <strong>of</strong> <strong>the</strong> transcription factors such as CRE<br />

or NF-κB are very downstream <strong>of</strong> GPCR activation (Siehler 2008), we, also measured [ 3H]inositol<br />

112 | P a g e


The role <strong>of</strong> isoleucine at position 6.45 in <strong>the</strong> H1R<br />

phosphate formation produced by <strong>the</strong> wild type H1R and <strong>the</strong> I6.45R H1R mutant expressed in COS-7<br />

as previously described (Bakker, Weiner et al. 2004).<br />

Table 3<br />

Pharmacological values <strong>of</strong> <strong>the</strong> wild type H1R and I6.45 H1R mutants<br />

Constructs<br />

NF-κB reporter gene assay<br />

Histamine Mepyramine Doxepin Ketotifen<br />

pEC50 %Effect pIC50 %Effect pIC50 %Effect pIC50 %Effect<br />

WT 6.5±0.1(4) +435±17(4) 7.9±0.1(3) -76±3(3) 8.3±0.1(3) -53±10 (3) 8.9±0.2(4) -47±10 (4)<br />

I 6.45 R c.n.d. c.n.d. c.n.d. c.n.d. 7.7±0.2(7)* +339±70(7) 9.0±0.1(3)* +510±118(3)<br />

I 6.45 N 6.8±0.2(3) +470±60 (3) 7.3±0.2 (3) -79±2 (3) 8.2±0.2 (5) -62±4 (5) 9.2±0.3 (6) -66±8 (6)<br />

Constructs<br />

CRE reporter gene assay<br />

Histamine Mepyramine Doxepin Ketotifen<br />

pEC50 %Effect pIC50 %Effect pIC50 %Effect pIC50 %Effect<br />

WT 6.1±0.1 (5) +553±96 (5) 7.2±0.1 (8) -51±6 (8) 7.9±0.3 (4) -68±4 (4) 8.7±0.1 (2) -75±4 (2)<br />

I 6.45 R c.n.d. c.n.d. c.n.d. c.n.d. 7.7±0.3(3)* +200±55(3) 8.9±0.2(4)* +443±47(4)<br />

I 6.45 N 6.1±0.1(4) +544±87 (4) 7.3±0.1 (9) -60±6 (9) 7.9±0.1 (2) -67±12 (2) 8.4±0.2 (3) -72±8 (3)<br />

Constructs<br />

Inositol triphosphate accumulation assay<br />

Histamine Mepyramine Doxepin Ketotifen<br />

pEC50 %Effect pIC50 %Effect pIC50 %Effect pIC50 %Effect<br />

WT 6.6±0.2 (3) +490±85(3) 8.8±0.1 (2) -50±10(2) 9.0±0.2 (3) -45±5 (3) 8.9±0.1 (2) -42±6 (2)<br />

I 6.45 R 5.2±0.1 (3) +56±4 (3) c.n.d. c.n.d. 8.4±0.2 (3)* +174±24 (3) 8.6±0.2 (2)* +170±10 (2)<br />

The values obtained from reporter gene assay and [ 3H]inositol phosphate formation <strong>of</strong> <strong>the</strong> wild type H1R and<br />

I6.45R H1R mutant expressed in COS-7 cells. Data shown are mean values ± S.E.M. (n, number <strong>of</strong><br />

experiments). c.n.d. = <strong>the</strong> value cannot be determined, * = pEC50<br />

The [ 3H]inositol phosphate formation assay relies on <strong>the</strong> incorporation <strong>of</strong> [ 3H]myo-inositol into<br />

cellular phosphatidyl-inositol 4,5-bisphosphate. Upon agonist binding to <strong>the</strong> H1R, <strong>the</strong> labeled<br />

phosphatidyl-inositol 4,5 bisphosphate is hydrolyzed by PLC to inositol-1,4,5-triphosphate (IP3). In<br />

<strong>the</strong> present <strong>of</strong> Li +, de-phosphorylation <strong>of</strong> IP3 to inositol is blocked, and <strong>the</strong> mass <strong>of</strong> soluble inositol<br />

phosphates is a quantitative readout <strong>of</strong> <strong>the</strong> H1R activation. As shown in Table 3 and Figure 5,<br />

histamine activates <strong>the</strong> wild type H1R (pEC50 6.6±0.1, N=3) and inverse agonists such as<br />

mepyramine, doxepin and ketotifen can inhibit <strong>the</strong> basal activity <strong>of</strong> <strong>the</strong> H1R. In contrast to <strong>the</strong> wild<br />

113 | P a g e


Chapter 6<br />

type H1R, <strong>the</strong> I6.45R H1R mutant showed a strongly decreased potency for histamine (pEC50<br />

5.3±0.1, N=3). The inverse agonists with a tricyclic moiety, such as doxepin and ketotifen, showed<br />

an agonistic effect on <strong>the</strong> I6.45R H1R mutant. These results are in a good agreement to that obtained<br />

from <strong>the</strong> reporter gene assays.<br />

Figure 5 [ 3H]inositol phosphate formation <strong>of</strong> <strong>the</strong> wild type H1R or <strong>the</strong> I6.45R H1R<br />

The [ 3 H]inositol phosphate formation assays were performed in COS-7 cells expressing with <strong>the</strong> wild type<br />

H1R or <strong>the</strong> I6.45R H1R mutated. The transfected cells were stimulated with H1R ligands such as histamine (�),<br />

mepyramine ( ), doxepin (•) and ketotifen (ο). Basal activity <strong>of</strong> each construct is put at 100%. Data shown<br />

are from a representative experiments (N≥ 2) performed in triplicate.<br />

The agonistic effects <strong>of</strong> <strong>the</strong> tricyclic H1R inverse agonists on <strong>the</strong> I6.45R H1R mutant can be explained<br />

using <strong>the</strong> β2AR-based model <strong>of</strong> <strong>the</strong> H1R (collaboration with Dr. L. Pardo). The used H1R model<br />

explains <strong>the</strong> role <strong>of</strong> several residues that are known to be involved in <strong>the</strong> binding pocket <strong>of</strong> <strong>the</strong> H1R<br />

such as D3.32 and W6.48 (Jongejan, Bruysters et al. 2005; Jongejan and Leurs 2005). In preliminary<br />

molecular dynamic simulation runs, <strong>the</strong> tricyclic rings <strong>of</strong> ketotifen binds into <strong>the</strong> aromatic pocket <strong>of</strong><br />

<strong>the</strong> wild type H1R, formed by F5.47, F6.52, and W6.48 and <strong>the</strong>reby inhibit <strong>the</strong> conformational<br />

transition <strong>of</strong> <strong>the</strong> W6.48 towards <strong>the</strong> active state (Figure 6) (see Chapter 1 for fur<strong>the</strong>r reading on<br />

<strong>the</strong> mechanism <strong>of</strong> receptor activation).<br />

Mutation <strong>of</strong> <strong>the</strong> isoleucine 6.45 into arginine adds a positive charge in <strong>the</strong> binding pocket, <strong>the</strong>reby<br />

inducing <strong>the</strong> polar atom <strong>of</strong> ketotifen to point towards <strong>the</strong> intracellular side and forming a hydrogen<br />

114 | P a g e


The role <strong>of</strong> isoleucine at position 6.45 in <strong>the</strong> H1R<br />

bond with W6.48. This network triggers <strong>the</strong> conformational transition <strong>of</strong> <strong>the</strong> W6.48 towards <strong>the</strong><br />

active state and leads to <strong>the</strong> activation <strong>of</strong> I6.45R H1R mutant. More experiments using homology<br />

modeling and/or fur<strong>the</strong>r mutation analysis need to be performed in order to clarify <strong>the</strong> role <strong>of</strong> I6.45<br />

in <strong>the</strong> TM6 <strong>of</strong> <strong>the</strong> H1R in <strong>the</strong> receptor activation mechanisms and to explain why <strong>the</strong> tricyclic H1R<br />

inverse agonists act as agonists at <strong>the</strong> I6.45R H1R mutant.<br />

Figure 6 Binding <strong>of</strong> ketotifen to <strong>the</strong> wild type H1R and <strong>the</strong> I6.45R H1R mutant<br />

The model showing <strong>the</strong> binding <strong>of</strong> ketotifen to <strong>the</strong> wild type H1R and <strong>the</strong> I6.45R H1R mutant was obtained<br />

from Dr. L. Pardo. The H1R model was based on <strong>the</strong> crystal structure <strong>of</strong> <strong>the</strong> β2AR. The tricyclic rings <strong>of</strong><br />

ketotifen bind inside <strong>the</strong> aromatic pocket <strong>of</strong> <strong>the</strong> wild type H1R. The polar atom <strong>of</strong> ketotifen forms a hydrogen<br />

bond with K5.39 and constrains <strong>the</strong> conformation <strong>of</strong> W6.48 in <strong>the</strong> inactive state. Mutation <strong>of</strong> <strong>the</strong> I6.45<br />

creates <strong>the</strong> positive charge inside <strong>the</strong> binding pocket. The polar atom <strong>of</strong> ketotifen is forced to form a<br />

hydrogen bond with <strong>the</strong> R6.45 residue and W6.48, thus stabilizing <strong>the</strong> active conformation <strong>of</strong> <strong>the</strong> receptor.<br />

Besides <strong>the</strong> I6.45R H1R mutant, we also characterized o<strong>the</strong>r H1R mutants at this residue.<br />

Interestingly, we have identified a high degree <strong>of</strong> constitutive activation from <strong>the</strong> mutation <strong>of</strong><br />

isoleucine to asparagine (I6.45N) (Figure 3). However, besides showing its high constitutive<br />

activity, <strong>the</strong> I6.45N H1R mutant does not possess a unique pharmacological pr<strong>of</strong>ile like <strong>the</strong> I6.45R<br />

H1R mutant. Histamine, mepyramine, doxepin and ketotifen modulate <strong>the</strong> I6.45N H1R mutant in a<br />

similar manner as <strong>the</strong> wild type H1R as determined by NFκB and CRE reporter gene assay (Table<br />

2). The mechanism <strong>of</strong> constitutive activation is, however, not yet known. Possibly, <strong>the</strong> I6.45N H1R<br />

115 | P a g e


Chapter 6<br />

mutant disrupts a bond between F6.44 and <strong>the</strong> TM 3 <strong>of</strong> <strong>the</strong> H1R, which was previously shown to be<br />

involved in <strong>the</strong> activation process in rhodopsin and o<strong>the</strong>r biogenic amines (Donnelly, Findlay et al.<br />

1994; Spalding, Burstein et al. 1998; Baranski, Herzmark et al. 1999; Chen, Lin et al. 2002), and<br />

stabilizes <strong>the</strong> I6.45N H1R in its active configuration. These findings, <strong>the</strong>refore, would fur<strong>the</strong>r<br />

underscore <strong>the</strong> importance <strong>of</strong> <strong>the</strong> isoleucine residue at position 6.45 in <strong>the</strong> H1R activation.<br />

116 | P a g e<br />

Conclusion<br />

In conclusion, we demonstrate that mutations at a single residue (I6.45) can induce multiple<br />

receptor species such as a high constitutive active mutant (I6.45N) and, more important, an<br />

exceptional mutant that does not respond to <strong>the</strong> endogenous ligand histamine (I6.45R) in reporter<br />

gene assays (NF-κB and CRE) and inositol triphosphate accumulation, but could be activated by<br />

only a particular class <strong>of</strong> syn<strong>the</strong>tic H1R inverse agonists that contain tricyclic moieties such as<br />

doxepin and ketotifen. Moreover, <strong>the</strong> I6.45R mutant shows no binding activity to mepyramine,<br />

while it still retains high affinity for <strong>the</strong> two o<strong>the</strong>r tricyclic inverse agonists; doxepin and ketotifen.<br />

This supports <strong>the</strong> previous finding (Nonaka, Otaki et al. 1998) that <strong>the</strong> biaryl H1R inverse agonists<br />

most likely bind differently compared to <strong>the</strong> tricyclic H1R inverse agonists. These pharmacological<br />

data, in combination with future molecular modeling studies will provide better insight on H1R<br />

ligand binding sites and <strong>the</strong> receptor activation process.


CHAPTER<br />

General discussion<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Chapter 7<br />

118 | P a g e<br />

General discussion<br />

G protein-coupled receptors (GPCRs) consist <strong>of</strong> a single polypeptide chain <strong>of</strong> 400-500 residues that<br />

wind seven times back and forth through <strong>the</strong> lipid bilayer connecting an extracellular amino N-<br />

terminus to an intracellular carboxy C-terminus. They are <strong>the</strong> largest class <strong>of</strong> membrane protein in<br />

<strong>the</strong> human genome and mediate responses from extracellular signals as diverse as biogenic amines,<br />

lipids, peptides, proteins, odorants, pheromones, tastants, ions, and photons. GPCRs are <strong>the</strong>refore<br />

involved in various physiological processes like smell, taste, vision, behavior, mood, regulation <strong>of</strong><br />

<strong>the</strong> immune system and autonomic nervous system transmission, which making <strong>the</strong>m an attractive<br />

target for <strong>the</strong> pharmaceutical industry (Jacoby, Bouhelal et al. 2006; Schlyer and Horuk 2006;<br />

Lefkowitz 2007). As shown in Table 1, six out <strong>of</strong> top ten best selling drugs are targeted to GPCRs.<br />

Table 1.<br />

Examples <strong>of</strong> top selling drugs (a)<br />

Molecular<br />

Entity<br />

Trade<br />

name<br />

Target Therapeutic<br />

Indication<br />

Atorvastatin Lipitor® Enzyme Hypercholesterolemia 13.5<br />

Clopidogrel Plavix® GPCR (P2Y12) Stroke 7.3<br />

Esomeprazole Nexium® Enzyme Dyspepsia 7.2<br />

Salmeterol Seretide® GPCR (β2) Asthma 7.1<br />

Etanercept Enbrel® TNFα Autoimmune diseases 5.3<br />

Olanzapine Zyprexa® GPCR(5HT2/D1/D2) Schizophrenia 5.0<br />

Risperidone Risperdal® GPCR (5HT2/D2) Schizophrenia 4.9<br />

Quetiapine Seroquel® GPCR (5HT2/D1/D2) Schizophrenia 4.6<br />

Montelukast Singulair® GPCR (LT) Asthma 4.5<br />

Darbepoetin Aranesp® Erythropoiesis Anemia 4.4<br />

World Sales<br />

(US$ billions)<br />

(a) Source: IMS MIDAS®, Dec 2007; Top ten global products 2007; All information current as <strong>of</strong> Feb 26, 2008;<br />

The report excludes unaudited markets, and Russia, Ukraine and Belarus.<br />

Abbreviation: P2Y12, purinergic receptor; β2, adrenergic receptor; 5HT2, hydroxytryptamine receptor; D1,D2,<br />

dopamine receptor; LT, leukotriene receptor<br />

Histamine H1 receptors (H1Rs) belong to <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs. They are expressed<br />

ubiquitously in smooth muscle, vascular endo<strong>the</strong>lium, gastro-intestinal tract, heart, and brain. Upon<br />

histamine binding, histamine exerts many actions related to allergic inflammation (Thurmond,<br />

Gelfand et al. 2008) in <strong>the</strong> skin (e.g. local reddening, wheal and flare), nose (e.g. nasal congestion<br />

and runny nose) and lower airways (e.g. bronchoconstriction). Fur<strong>the</strong>rmore, <strong>the</strong> H1R is also


General discussion<br />

responsible for many <strong>of</strong> <strong>the</strong> CNS effect <strong>of</strong> histamine, such as <strong>the</strong> reduction <strong>of</strong> locomotor activity,<br />

exploratory behavior, and cognitive functions (Inoue, Yanai et al. 1996). Knowledge regarding 3D<br />

structural detail will facilitate <strong>the</strong> understanding <strong>of</strong> how <strong>the</strong> H1R and <strong>the</strong> o<strong>the</strong>r GPCRs function at<br />

atomic level (Pierce, Premont et al. 2002; Horn, Bettler et al. 2003; Akermoun, Koglin et al. 2005;<br />

Lundstrom 2005). However, due to <strong>the</strong> low expression in native tissues and <strong>the</strong> difficulties in GPCR<br />

purification and crystallization (Sarramegna, Talmont et al. 2003; Lundstrom, Wagner et al. 2006),<br />

<strong>the</strong> structural information <strong>of</strong> <strong>the</strong>se important membrane proteins has been limited to crystal<br />

structures <strong>of</strong> <strong>the</strong> naturally abundant GPCR, rhodopsin [31,54,56,71], <strong>the</strong> β-adrenergic receptors<br />

[40-42] and <strong>the</strong> A2A adenosine receptor(Jaakola, Griffith et al. 2008). The aim <strong>of</strong> <strong>the</strong> research<br />

described in this PhD <strong>the</strong>sis is to gain insight in structural features <strong>of</strong> <strong>the</strong> H1R using techniques such<br />

as mass spectrometry and site-directed mutagenesis in combination with homology modeling<br />

studies. Using a new in vitro cell free expression systems, we were able to express milligrams <strong>of</strong> <strong>the</strong><br />

H1R protein within 8-12 hours. The protein obtained can be purified by a simple and cost effective<br />

IMAC chromatography and reconstituted into artificial membranes prepared from asolectin lipids.<br />

The purified H1R was analyzed by ligand binding analysis using [ 3H]mepyramine as a radiotracer.<br />

The cell free-produced H1R protein retains its binding activity towards <strong>the</strong> inverse agonist<br />

mepyramine and its endogenous ligand histamine and does not bind to H2, H3 and H4 ligands<br />

(Chapter 3). The rapid generation <strong>of</strong> sufficient amount <strong>of</strong> <strong>the</strong> H1R protein using this in vitro cell free<br />

expression system allowed us to use <strong>the</strong> H1R protein in mass spectrometry (MS) analysis. With its<br />

high sensitivity, toge<strong>the</strong>r with continuous improvements in technology and s<strong>of</strong>tware, MS<br />

technology has become an important tool in <strong>the</strong> sequencing and characterizing <strong>of</strong> proteins and<br />

protein complex. By employing an acid labile surfactant during <strong>the</strong> tryptic digestion prior to<br />

MALDI-ToF MS analysis, we were able to obtain high sequence coverage <strong>of</strong> <strong>the</strong> H1R, including five<br />

out <strong>of</strong> seven transmembrane (TM) domains (Chapter 3). Following <strong>the</strong> same approach for protein<br />

production using an in vitro cell free expression system using MALDI-ToF MS, we were able to<br />

produce milligrams amount <strong>of</strong> <strong>the</strong> H4R, a new potential target for drug discovery. Coverage <strong>of</strong><br />

~50% amino acid sequence <strong>of</strong> <strong>the</strong> H4R, including three out <strong>of</strong> seven TM domains, was obtained<br />

(Chapter 3). Future works should focus on <strong>the</strong> improvement <strong>of</strong> <strong>the</strong> percent coverage <strong>of</strong> <strong>the</strong> amino<br />

acid sequence and <strong>the</strong> functional characterization <strong>of</strong> <strong>the</strong> purified and reconstituted H4R protein.<br />

Employing a similar mass spectrometric analysis approach, we were also able to obtain <strong>the</strong> high<br />

sequence coverage <strong>of</strong> <strong>the</strong> H1R protein expressed from baculovirus/insect cells expression system as<br />

previously described by Ratnala and co-workers [126]. Using MS analysis, we show here for <strong>the</strong><br />

first time that an asparagine residue in <strong>the</strong> H1R protein is subjected to <strong>the</strong> N-linked glycosylation<br />

119 | P a g e


Chapter 7<br />

(Chapter 4). These data suggests that MS technology is a promising tool for <strong>the</strong> characterization <strong>of</strong><br />

higher order protein structures, post-translation modifications, and receptor-ligand binding<br />

interactions.<br />

In Chapter 4, we also report on a preliminary MS experiment in map residues that involved in <strong>the</strong><br />

H1R-ligand binding. The ligand phenoxybenzamine was found to irreversibly bind to <strong>the</strong> H1R,<br />

expressed by baculovirus/insect cells expression system and <strong>the</strong> cysteine residue (Cys302) located<br />

in <strong>the</strong> intracellular loop <strong>of</strong> <strong>the</strong> H1R was identified by MS analysis to be alkylated by<br />

phenoxybenzamine. However, based on <strong>the</strong> structured model <strong>of</strong> <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs,<br />

including <strong>the</strong> H1R, <strong>the</strong> binding pocket is thought to reside within <strong>the</strong> TM domains, such as TM3, TM5<br />

and TM6 [See Chapter 1 for fur<strong>the</strong>r reading]. Fur<strong>the</strong>rmore, it has also been shown that <strong>the</strong> possible<br />

cysteine reside that is involved in <strong>the</strong> binding <strong>of</strong> phenoxybenzamine is located at position 3.36 <strong>of</strong><br />

TM3 in <strong>the</strong> α-adrenergic receptor [256]. Therefore, we speculate that our preliminary finding that<br />

we obtained from mapping <strong>of</strong> phenoxybenzamine and <strong>the</strong> H1R (at <strong>the</strong> Cys302) is <strong>the</strong> result from <strong>the</strong><br />

non-specific reaction <strong>of</strong> phenoxybenzamine [314]. Recently, <strong>the</strong> works <strong>of</strong> Espinoza-Fonseca [312]<br />

and Nicholls [313] have, however, suggested possible binding sites <strong>of</strong> GPCR ligands to <strong>the</strong><br />

intracellular region. It is <strong>the</strong>refore interesting to fur<strong>the</strong>r investigate <strong>the</strong> binding site <strong>of</strong><br />

phenoxybenzamine and <strong>the</strong> H1R. For future MS analysis, <strong>the</strong> protocol regarding <strong>the</strong><br />

phenoxybenzamine treatment and <strong>the</strong> H1R sample preparation can be adjusted to minimize a<br />

potential non-specific reaction. It is also noteworthy to study <strong>the</strong> intracellular region <strong>of</strong> <strong>the</strong> H1R as a<br />

possible binding site by mutation analysis toge<strong>the</strong>r with homology modeling studies.<br />

In Chapter 5, we have combined <strong>the</strong> latest structural insights in GPCR structures, computation<br />

modeling and site-directed mutagenesis <strong>of</strong> <strong>the</strong> H1R to study <strong>the</strong> role <strong>of</strong> <strong>the</strong> TAS(I/V) motif at<br />

positions 3.37-3.40 in TM3 in <strong>the</strong> process <strong>of</strong> receptor activation. We have designed point mutations<br />

that modulate <strong>the</strong> strength <strong>of</strong> intra-molecular interactions and <strong>the</strong> side chain rotamer distribution<br />

<strong>of</strong> specific residues, and measured <strong>the</strong>ir effects on constitutive agonist-independent H1R activity.<br />

We have shown here that <strong>the</strong> threonine T3.37 interacts with TM5 in <strong>the</strong> inactive state <strong>of</strong> <strong>the</strong> H1R,<br />

while <strong>the</strong> serine S3.39 changes its conformation upon receptor activation from <strong>the</strong> inactive g+ to<br />

<strong>the</strong> active g- or t conformer, toge<strong>the</strong>r with S3.36, to achieve new water-mediated inter-helical<br />

interactions with TM2 and TM7 that stabilize <strong>the</strong> active state. We also provided <strong>the</strong> first evidence<br />

that <strong>the</strong> hydrophobic side chain at position 3.40, which is highly conserved in <strong>the</strong> rhodopsin family<br />

<strong>of</strong> GPCRs, plays a key role in facilitating <strong>the</strong> histamine-induced relocation <strong>of</strong> TM5. Mutation <strong>of</strong> I3.40<br />

120 | P a g e


General discussion<br />

to ei<strong>the</strong>r Ala or Gly abolished <strong>the</strong> constitutive activity <strong>of</strong> <strong>the</strong> H1R and well as <strong>the</strong> histamine-induced<br />

receptor activation. A thorough understanding on how a ligand induced <strong>the</strong> receptor will have a big<br />

impact on our understanding <strong>of</strong> <strong>the</strong> molecular aspects <strong>of</strong> <strong>the</strong> H1R and <strong>the</strong> related family<br />

A GPCRs. In 2008, Bakker and co-workers [62] have employed random mutagenesis studies and<br />

identified three residues in TM6 (I6.40, A6.43 and I6.45) as sites prone to yield high constitutively<br />

active H1R mutants. Following <strong>the</strong>ir work, we have found <strong>the</strong> I6.45R H1R mutant to possess a<br />

unique pharmacological pr<strong>of</strong>ile (Chapter 6). The I6.45R H1R mutant loses its binding activity<br />

towards <strong>the</strong> classical biaryl H1R inverse agonist mepyramine, but still retains its affinity towards<br />

tricyclic H1R inverse agonists such as doxepin and ketotifen, supporting <strong>the</strong> previous finding by<br />

Nonaka and co-workers[253] that <strong>the</strong> H1R inverse agonists possibly bind differently to <strong>the</strong> H1R. The<br />

I6.45R H1R mutant also loses its basal activity and cannot be fur<strong>the</strong>r activated by histamine as<br />

measured by reporter gene and [ 3H]inositol formation assays, despite its high affinity towards<br />

histamine. Surprisingly, <strong>the</strong> mutant can only be activated by doxepin and ketotifen, <strong>the</strong> tricyclic H1R<br />

inverse agonists. Here, we propose a possible hypo<strong>the</strong>sis why <strong>the</strong> I6.45R H1R mutant is activated by<br />

ketotifen using a H1R model based on <strong>the</strong> crystal structure <strong>of</strong> <strong>the</strong> β2 adrenergic receptor. The<br />

arginine R6.45 in <strong>the</strong> H1R mutant provides a positive charge in <strong>the</strong> binding pocket, <strong>the</strong>reby inducing<br />

<strong>the</strong> polar atom <strong>of</strong> tricyclic inverse agonists such as ketotifen to point towards <strong>the</strong> intracellular side<br />

and to trigger <strong>the</strong> conformational transition <strong>of</strong> <strong>the</strong> W6.48 towards <strong>the</strong> active state via <strong>the</strong> hydrogen<br />

bond network.<br />

121 | P a g e


Summary<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Summary<br />

124 | P a g e<br />

Summary<br />

Histamine is a biogenic amine well-known for its role in allergic reactions. When histamine is<br />

released in <strong>the</strong> body, for example, after specific antigens, such as pollen or dust, come in contact<br />

with mast cells, it exerts its effect by binding to histamine receptors. Histamine receptors are G<br />

protein-coupled receptors (GPCRs) or seven-transmembrane domain (7TM) receptors, which<br />

represent <strong>the</strong> largest class <strong>of</strong> membrane-bound proteins in <strong>the</strong> human genome. Some <strong>of</strong> <strong>the</strong> best-<br />

selling drugs on <strong>the</strong> market today such as Allegra® and Claritin® target histamine receptors. So far<br />

four subtypes <strong>of</strong> histamine receptors have been discovered and cloned, and <strong>the</strong> research described<br />

in this <strong>the</strong>sis is mainly focused on <strong>the</strong> histamine H1 receptor (H1R). Even before <strong>the</strong> cloning <strong>of</strong> <strong>the</strong><br />

human H1R, scientists have been able to develop antagonists that potently block histamine from<br />

binding to <strong>the</strong> H1R. Through <strong>the</strong> years a steady increase in <strong>the</strong> structural knowledge <strong>of</strong> <strong>the</strong><br />

receptor, has led to increases in selectivity and strength <strong>of</strong> <strong>the</strong>se antagonists, while at <strong>the</strong> same time<br />

led to a reduction in side effects such as sedation. Pharmaceutical companies rely heavily on<br />

structure-based design for lead optimization in drug discovery, especially if X-ray crystal structures<br />

<strong>of</strong> a target protein are available. It is widely used in drug discovery for soluble proteins, because<br />

<strong>the</strong>y are more straightforward to isolate, purify and crystallize in comparisons to membrane<br />

proteins. Despite <strong>the</strong> importance <strong>of</strong> membrane proteins such as GPCRs as drug targets, <strong>the</strong>ir<br />

structure determination has been very limited compared to soluble proteins. The low success rate<br />

is due to <strong>the</strong> difficulties in obtaining large amounts <strong>of</strong> high quality proteins. The research described<br />

in this <strong>the</strong>sis was aimed to provide a better understanding <strong>of</strong> GPCR structure in general and <strong>of</strong> <strong>the</strong><br />

human H1R in specific, through large scale production <strong>of</strong> <strong>the</strong> H1R and subsequent purification and<br />

structural analysis by mass spectrometry. Chapter 1 gives a brief overview <strong>of</strong> <strong>the</strong> current<br />

knowledge <strong>of</strong> GPCRs and H1R in particular. Chapter 2 provides a systematic and thorough review <strong>of</strong><br />

<strong>the</strong> different options available to express and purify GPCRs, including <strong>the</strong> H1R protein. In Chapter 3,<br />

we described how we have made use <strong>of</strong> a well established baculovirus/insect cell expression<br />

system, and an in vitro cell free expression system to produce large amounts <strong>of</strong> <strong>the</strong> human H1R and<br />

H4R protein. Moreover, this chapter also shows details in <strong>the</strong> use <strong>of</strong> immobilized metal affinity<br />

chromatography (IMAC) to purify <strong>the</strong> receptors after production, as well as our endeavors to obtain<br />

<strong>the</strong> highest possible sequence coverage <strong>of</strong> <strong>the</strong> two receptors by MALDI-ToF and LC-MS/MS. Having<br />

established a successful and reproducible method to obtain large and pure enough quantities <strong>of</strong> <strong>the</strong><br />

human H1R protein allowed us to continue our goal <strong>of</strong> improving <strong>the</strong> structural knowledge and<br />

binding pocket <strong>of</strong> <strong>the</strong> H1R. In chapter 4, we describe how we employed MS analysis to discover that<br />

an asparagine residue in <strong>the</strong> N-terminus <strong>of</strong> <strong>the</strong> H1R (Asn5) is glycosylated in insect cells, which had


Summary<br />

been used in conjunction with <strong>the</strong> baculovirus/insect cell expression system. Moreover we were<br />

able to locate <strong>the</strong> residue to which phenoxybenzamine, an irreversible H1R antagonist, covalently<br />

binds to <strong>the</strong> H1R by using MS analysis, <strong>the</strong>reby showing <strong>the</strong> great potential <strong>of</strong> our methods in <strong>the</strong><br />

production, purification and MS analysis <strong>of</strong> <strong>the</strong> H1R protein could have on structural analysis <strong>of</strong><br />

GPCRs. It is known that GPCRs can become activated in <strong>the</strong> absence <strong>of</strong> ligands. The degree <strong>of</strong> this<br />

constitutive activity can also be modulated by single point mutations <strong>of</strong> <strong>the</strong> receptor. Constitutive<br />

activity <strong>of</strong> receptors can be used to investigate which residues in a GPCR contributes to receptor<br />

activation and how. In chapter 5 we used site directed mutagensis <strong>of</strong> <strong>the</strong> H1R to identify residues<br />

that are involved in activation <strong>of</strong> <strong>the</strong> receptor. Our results suggest that T3.37 interacts with TM5 in<br />

<strong>the</strong> inactive state <strong>of</strong> <strong>the</strong> receptor, while S3.39 changes conformation upon receptor activation to<br />

achieve, toge<strong>the</strong>r with S3.36, new water-mediated inter-helical interactions with TM2 and TM7 that<br />

stabilize <strong>the</strong> active state. In addition, we provide <strong>the</strong> first evidence that <strong>the</strong> hydrophobic side chain<br />

at position 3.40, highly conserved in <strong>the</strong> rhodopsin family <strong>of</strong> GPCRs, plays a key role in facilitating<br />

<strong>the</strong> agonist-induced relocation <strong>of</strong> TM5. Mutation <strong>of</strong> I3.40 to ei<strong>the</strong>r alanine or glycine, i.e. removing<br />

<strong>the</strong> bulky side chain at this position, abolishes <strong>the</strong> constitutive activity <strong>of</strong> <strong>the</strong> H1R, <strong>the</strong> effect <strong>of</strong><br />

constitutive-activity increasing mutations, as well as <strong>the</strong> histamine-induced receptor activation. In<br />

chapter 6 we describe our detailed investigation <strong>of</strong> a point mutated version <strong>of</strong> <strong>the</strong> H1R. Previous<br />

research had identified an isoleucine residue at position 6.45 <strong>of</strong> <strong>the</strong> H1R to be prone to yield high<br />

constitutively active mutants. We observed that <strong>the</strong> I6.45R mutant exhibits an exceptional<br />

pharmacological pr<strong>of</strong>ile as it no longer gives a functional response when stimulated with <strong>the</strong><br />

endogenous ligand histamine, but only responds to a particular class <strong>of</strong> syn<strong>the</strong>tic H1R inverse<br />

agonists that contain tricyclic moieties such as doxepin and ketotifen. Mepyramine, an inverse<br />

agonist belonging to <strong>the</strong> class <strong>of</strong> biaryl H1R inverse agonists, is also unable to bind to <strong>the</strong> I6.45R<br />

mutant, suggesting that <strong>the</strong>se two classes <strong>of</strong> H1R inverse agonists may bind <strong>the</strong> H1R in a different<br />

manner.<br />

125 | P a g e


Samenvatting<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Samenvatting<br />

128 | P a g e<br />

Samenvatting<br />

Histamine is een amine, dat door het lichaam wordt geproduceerd. Het is voornamelijk bekend voor<br />

zijn grote rol in allergische reacties. Wanneer histamine vrijkomt, bijvoorbeeld nadat een antigen,<br />

zoals pollen, in contact komt met een mastcel, kan de amine binden aan histaminereceptoren en zo<br />

een effect in de cel bewerkstelligen. Histaminereceptoren zijn G-eiwit gekoppelde receptoren<br />

(GPCRs) <strong>of</strong> zeven transmembraan domein (7TM) receptoren. Deze receptoren vormen de grootste<br />

klasse van celmembraangebonden eiwitten. Verscheidene “blockbuster” medicijnen richten zich op<br />

deze receptoren, inclusief een paar die zich specifiek op de histamine H1 receptor (H1R) richten<br />

(Allegra, Claritin). Tot dusver zijn er vier verschillende histaminereceptoren ontdekt en gekloond.<br />

Het onderzoek dat voor dit proefschrift is uitgevoerd is echter voornamelijk gericht op de H1R . Al<br />

lang voordat de H1R gekloond was, was het wetenschappers gelukt om antagonisten, die de binding<br />

van histamine aan de H1R kunnen voorkomen, te creëren. Over de jaren is de selectiviteit en<br />

affiniteit van deze antagonisten toegenomen en tegelijkertijd zijn vervelende bijwerkingen als<br />

slaperigheid afgenomen door een immer groeiend inzicht in de structuur van de H1R.<br />

Pharmaceutische bedrijven zijn erg afhankelijk van de kennis van de structuur van de receptor voor<br />

de ontwikkeling en verbetering van geneesmiddelen. Het verkrijgen van een X-ray kristalstructuur<br />

is erg informatief voor dit doeleinde, en dit wordt dan ook veel toegepast, met name voor eiwitten<br />

die gemakkelijk in oplossing blijven. Deze eiwitten zijn in tegenstelling tot GPCRs makkelijker te<br />

produceren en op te zuiveren en uiteindelijk te kristalliseren. Ondanks de grote interesse in GPCRs<br />

voor de ontwikkeling van geneesmiddelen is het vooralsnog moeilijk om deze receptor-eiwitten te<br />

kristalliseren <strong>of</strong> überhaupt goede structuur informatie te vergaren. Een van de redenen hiervoor is<br />

dat het lastig is om deze receptoren in grote hoeveelheden te produceren. Het onderzoek dat wordt<br />

beschreven in dit proefschrift was gericht op het verkrijgen van een beter inzicht de structuur van<br />

GPCRs en specifiek de structuur van de H1R, door middel van het produceren van de H1R op grote<br />

schaal en het vervolgens analyseren van de gezuiverde receptor via massaspectrometrie. Ho<strong>of</strong>dstuk<br />

1 bevat een kort overzicht van de huidige kennis op het gebied van GPCRs. Ho<strong>of</strong>dstuk 2 geeft een<br />

uitgebreid en systematisch overzicht van alle opties en methoden die beschikbaar zijn voor het<br />

produceren en zuiveren van grote hoeveelheden GPCRs. In dit proefschrift hebben we verschillende<br />

methoden gebruikt om grote hoeveelheden H1R te verkrijgen. We hebben gebruik gemaakt van het<br />

baculovirus systeem, wat een veel gebruikte methode is voor dit doeleinde. We hebben echter ook<br />

een relatief nieuwe methode gebruikt. Deze methode is gebaseerd op een cell-vrij<br />

expressiesysteem. In ho<strong>of</strong>dstuk 3 beschrijven we hoe we het baculovirus en het cell-vrije<br />

expressiesysteem hebben gebruikt om H1R en H4R te produceren. Verder beschrijven we hoe we de


Samenvatting<br />

receptors hebben gezuiverd via een geimmobiliseerde metaal-affiniteitschromotografie en onze<br />

pogingen om de hoogste dekking te verkrijgen in de structuuranalyse van de receptoren met<br />

behulp van MALDI-TOF en LC MS/MS.<br />

Nadat het ons gelukt was om een succesvolle en reproduceerbare methode op te zetten voor de<br />

productie van H1Rs, konden we ons richten op ons doel een beter inzicht in de structuur van de H1R<br />

te verkrijgen. In ho<strong>of</strong>dstuk 4 beschrijven we hoe we massaspectrometrie hebben gebruikt om te<br />

ontdekken dat een asparagine in de amino-terminale staart van de H1R (Asn5) is geglycolyseerd in<br />

insectencellen. Het was ook mogelijk om een aminozuur, dat betrokken is bij de onomkeerbare<br />

binding van de H1R antagonist phenoxybenzamine aan de receptor te lokaliseren. Hiermee konden<br />

we ook het grote potentieel van onze methode aantonen, voor het analyseren van de structuur van<br />

de bindingsplaats van geneesmiddelen aan de H1R en mogelijk ook andere GPCRs.<br />

Het is bekend dat GPCRs ook zonder de binding van een agonist in een geactiveerde staat kan<br />

bestaan. De sterkte van deze constitutieve activiteit kan gemanipuleerd worden met behulp van<br />

puntmutaties in de receptor. Constitutieve activiteit van receptoren kan worden gebruikt om de<br />

aminozuren, die een rol spelen bij de activering van de receptor, te bestuderen. Voor ho<strong>of</strong>dstuk 5<br />

hebben we gebruik gemaakt van zogenaamde ‘site-directed’ mutagenese om aminozuren in de H1R<br />

die betrokken zijn bij de activatie van de receptor te identificeren. Onze resultaten suggereren dat<br />

de T3.37 een interactie aangaat met TM5 in de inactieve staat van de receptor. Wanneer de<br />

receptor geactiveerd wordt, verandert de conformatie van S3.39, zodat dit residue samen met S3.36<br />

een water-geassisteerde interactie aangaat tussen TM3, TM2 en TM7. Deze interactie stabiliseert de<br />

actieve staat van de receptor. Tegelijkertijd leveren we ook het eerste bewijs dat de hydr<strong>of</strong>obische<br />

zijketting op positie 3.40 een grote rol speelt in het faciliteren van de herschikking van TM5 nadat<br />

de receptor is geactiveerd. Door het verwijderen van de grote zijketting rond 3.40 door mutaties<br />

van I3.40 naar een alanine <strong>of</strong> een glycine, wordt zowel de constitutieve- als ook de agonist-<br />

geïnduceerde acitiveit van de H1R tenietgedaan.<br />

In ho<strong>of</strong>dstuk 6 beschrijven we ons onderzoek van een specifieke puntmutatie in de H1R. Eerder<br />

onderzoek had een isoleucine op positie 6.45 geïdentificeerd, die erg vatbaar was om mutanten te<br />

creëren met hoge constitutieve activiteit. We observeerden dat de I6.45R-mutant een opmerkelijk<br />

farmacologisch pr<strong>of</strong>iel had. De mutant reageerde niet meer op de binding van histamine, maar<br />

enkel op binding van een specifieke klasse van syn<strong>the</strong>tische H1R inverse-agonisten met een<br />

tricyclisch karakter, zoals doxepine en ketotifen. Mepyramine, een andere klasse H1R inverse-<br />

129 | P a g e


Samenvatting<br />

agonist, kon echter niet aan deze mutant binden. Dit doet vermoeden dat deze twee klassen van<br />

inverse agonisten op een verschillende manier aan de H1R binden.<br />

130 | P a g e


References<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


References<br />

References<br />

Abdulaev, N. G. and K. D. Ridge (2000). "Heterologous expression <strong>of</strong> bovine opsin in Pichia pastoris." Methods<br />

Enzymol 315: 3-11.<br />

Ablonczy, Z., R. K. Crouch, et al. (2005). "Mass spectrometric analysis <strong>of</strong> integral membrane proteins at <strong>the</strong><br />

subpicomolar level: application to rhodopsin." J Chromatogr B Analyt Technol Biomed Life Sci 825(2): 169-<br />

75.<br />

Ablonczy, Z., P. Goletz, et al. (2002). "Mass spectrometric analysis <strong>of</strong> porcine rhodopsin." Photochem Photobiol 75(3):<br />

316-21.<br />

Ablonczy, Z., M. Kono, et al. (2001). "Mass spectrometric analysis <strong>of</strong> integral membrane proteins at <strong>the</strong> subnanomolar<br />

level: application to recombinant photopigments." Anal Chem 73(20): 4774-9.<br />

Ahuja, S., E. Crocker, et al. (2009). "Location <strong>of</strong> <strong>the</strong> retinal chromophore in <strong>the</strong> activated state <strong>of</strong> rhodopsin." J. Biol.<br />

Chem.<br />

Akdis, C. A. and F. E. Simons (2006). "Histamine receptors are hot in immunopharmacology." Eur J Pharmacol 533(1-<br />

3): 69-76.<br />

Akermoun, M., M. Koglin, et al. (2005). "Characterization <strong>of</strong> 16 human G protein-coupled receptors expressed in<br />

baculovirus-infected insect cells." Protein Expr Purif 44(1): 65-74.<br />

Akiba, S., Y. Hayashi, et al. (2006). "Extracellular production <strong>of</strong> human cystatin S and cystatin SA by Bacillus subtilis."<br />

Protein Expr Purif 49(2): 203-10.<br />

Akiba, T., C. Toyoshima, et al. (1996). "Three-dimensional structure <strong>of</strong> bovine cytochrome bc1 complex by electron<br />

cryomicroscopy and helical image reconstruction." Nat Struct Biol 3(6): 553-61.<br />

Alewijnse, A. E., H. Timmerman, et al. (2000). "The effect <strong>of</strong> mutations in <strong>the</strong> DRY motif on <strong>the</strong> constitutive activity and<br />

structural instability <strong>of</strong> <strong>the</strong> histamine H(2) receptor." Mol Pharmacol 57(5): 890-8.<br />

Alkhalfioui, F., Magnin, T, Wagner, R (2009). "From purified GPCRs to drug discovery: <strong>the</strong> promise <strong>of</strong> protein-based<br />

methodologies." COPHAR 9: 1-7.<br />

Almog, S., T. Kushnir, et al. (1986). "Kinetic and structural aspects <strong>of</strong> reconstitution <strong>of</strong> phosphatidylcholine vesicles by<br />

dilution <strong>of</strong> phosphatidylcholine-sodium cholate mixed micelles." Biochemistry 25(9): 2597-605.<br />

Altenbach, C., J. Klein-Seetharaman, et al. (2001). "Structure and function in rhodopsin: mapping light-dependent<br />

changes in distance between residue 316 in helix 8 and residues in <strong>the</strong> sequence 60-75, covering <strong>the</strong><br />

cytoplasmic end <strong>of</strong> helices TM1 and TM2 and <strong>the</strong>ir connection loop CL1." Biochemistry 40(51): 15493-500.<br />

Altmann, F., E. Staudacher, et al. (1999). "Insect cells as hosts for <strong>the</strong> expression <strong>of</strong> recombinant glycoproteins."<br />

Glycoconj J 16(2): 109-23.<br />

Alves, I. D., E. Sachon, et al. (2007). "Analysis <strong>of</strong> an Intact G-Protein Coupled Receptor by MALDI-TOF Mass<br />

Spectrometry: Molecular Heterogeneity <strong>of</strong> <strong>the</strong> Tachykinin NK-1 Receptor." Anal Chem 79(6): 2189-2198.<br />

Andersen, B. and R. C. Stevens (1998). "The human D1A dopamine receptor: heterologous expression in<br />

Saccharomyces cerevisiae and purification <strong>of</strong> <strong>the</strong> functional receptor." Protein Expr Purif 13(1): 111-9.<br />

Andre, N., N. Cherouati, et al. (2006). "Enhancing functional production <strong>of</strong> G protein-coupled receptors in Pichia<br />

pastoris to levels required for structural studies via a single expression screen." Protein Sci 15(5): 1115-26.<br />

Arkinstall, S., M. Edgerton, et al. (1995). "Co-expression <strong>of</strong> <strong>the</strong> neurokinin NK2 receptor and G-protein components in<br />

<strong>the</strong> fission yeast Schizosaccharomyces pombe." FEBS Lett 375(3): 183-7.<br />

Arnau, J., C. Lauritzen, et al. (2006). "Current strategies for <strong>the</strong> use <strong>of</strong> affinity tags and tag removal for <strong>the</strong> purification<br />

<strong>of</strong> recombinant proteins." Protein Expr Purif 48(1): 1-13.<br />

Asmann, Y. W., M. Dong, et al. (2004). "Functional characterization and purification <strong>of</strong> <strong>the</strong> secretin receptor expressed<br />

in baculovirus-infected insect cells." Regul Pept 123(1-3): 217-23.<br />

Audet, M. and M. Bouvier (2008). "Insights into signaling from <strong>the</strong> beta2-adrenergic receptor structure." Nat Chem<br />

Biol 4(7): 397-403.<br />

Bach, M., S. Grigat, et al. (2007). "Fast set-up <strong>of</strong> doxycycline-inducible protein expression in human cell lines with a<br />

single plasmid based on Epstein-Barr virus replication and <strong>the</strong> simple tetracycline repressor." Febs J 274(3):<br />

783-90.<br />

Bakker, R. A., G. Dees, et al. (2004). "Domain swapping in <strong>the</strong> human histamine H1 receptor." J Pharmacol Exp Ther<br />

311(1): 131-8.<br />

Bakker, R. A., A. Jongejan, et al. (2008). "Constitutively active mutants <strong>of</strong> <strong>the</strong> histamine H1 receptor suggest a<br />

conserved hydrophobic asparagine-cage that constrains <strong>the</strong> activation <strong>of</strong> class A G protein-coupled<br />

receptors." Mol Pharmacol 73(1): 94-103.<br />

132 | P a g e


References<br />

Bakker, R. A., S. B. Schoonus, et al. (2001). "Histamine H(1)-receptor activation <strong>of</strong> nuclear factor-kappa B: roles for G<br />

beta gamma- and G alpha(q/11)-subunits in constitutive and agonist-mediated signaling." Mol Pharmacol<br />

60(5): 1133-42.<br />

Bakker, R. A., D. M. Weiner, et al. (2004). "8R-lisuride is a potent stereospecific histamine H1-receptor partial agonist."<br />

Mol Pharmacol 65(3): 538-49.<br />

Baldi, L., D. L. Hacker, et al. (2007). "Recombinant protein production by large-scale transient gene expression in<br />

mammalian cells: state <strong>of</strong> <strong>the</strong> art and future perspectives." Biotechnol Lett 29(5): 677-84.<br />

Ball, L. E., J. E. Oatis, Jr., et al. (1998). "Mass spectrometric analysis <strong>of</strong> integral membrane proteins: application to<br />

complete mapping <strong>of</strong> bacteriorhodopsins and rhodopsin." Protein Sci 7(3): 758-64.<br />

Ballesteros, J. A., X. Deupi, et al. (2000). "Serine and Threonine Residues Bend a-Helices in <strong>the</strong> c1=g- conformation."<br />

Biophys. J. 79(5): 2754-2760.<br />

Ballesteros, J. A., A. D. Jensen, et al. (2001). "Activation <strong>of</strong> <strong>the</strong> beta 2-adrenergic receptor involves disruption <strong>of</strong> an ionic<br />

lock between <strong>the</strong> cytoplasmic ends <strong>of</strong> transmembrane segments 3 and 6." J Biol Chem 276(31): 29171-7.<br />

Ballesteros, J. A. and H. Weinstein (1995). "Integrated methods for <strong>the</strong> construction <strong>of</strong> three dimensional models and<br />

computational probing <strong>of</strong> structure-function relations in G-protein coupled receptors." Methods in<br />

Neurosciences 25: 366-428.<br />

Bane, S. E., J. E. Velasquez, et al. (2007). "Expression and purification <strong>of</strong> milligram levels <strong>of</strong> inactive G-protein coupled<br />

receptors in E. coli." Protein Expr Purif 52(2): 348-55.<br />

Baneres, J. L., A. Martin, et al. (2003). "Structure-based analysis <strong>of</strong> GPCR function: conformational adaptation <strong>of</strong> both<br />

agonist and receptor upon leukotriene B4 binding to recombinant BLT1." J Mol Biol 329(4): 801-14.<br />

Baneres, J. L., D. Mesnier, et al. (2005). "Molecular characterization <strong>of</strong> a purified 5-HT4 receptor: a structural basis for<br />

drug efficacy." J Biol Chem 280(21): 20253-60.<br />

Bansal, M., S. Kumar, et al. (2000). "HELANAL: a program to characterize helix geometry in proteins." J. Biomol. Struct.<br />

Dyn. 17(5): 811-9.<br />

Baranski, T. J., P. Herzmark, et al. (1999). "C5a receptor activation. Genetic identification <strong>of</strong> critical residues in four<br />

transmembrane helices." J Biol Chem 274(22): 15757-65.<br />

Barger, G. and H. H. Dale (1910). "Chemical structure and sympathomimetic action <strong>of</strong> amines." J Physiol 41(1-2): 19-<br />

59.<br />

Barnidge, D. R., E. A. Dratz, et al. (1999). "Extraction method for analysis <strong>of</strong> detergent-solubilized bacteriorhodopsin<br />

and hydrophobic peptides by electrospray ionization mass spectrometry." Anal Biochem 269(1): 1-9.<br />

Bartus, C. L., V. P. Jaakola, et al. (2003). "Downstream coding region determinants <strong>of</strong> bacterio-opsin, muscarinic<br />

acetylcholine receptor and adrenergic receptor expression in Halobacterium salinarum." Biochim Biophys<br />

Acta 1610(1): 109-23.<br />

Berridge, M. J. (1993). "Inositol trisphosphate and calcium signalling." Nature 361(6410): 315-25.<br />

Bertin, B., M. Freissmuth, et al. (1992). "Functional expression <strong>of</strong> <strong>the</strong> human serotonin 5-HT1A receptor in Escherichia<br />

coli. Ligand binding properties and interaction with recombinant G protein alpha-subunits." J Biol Chem<br />

267(12): 8200-6.<br />

Bhattacharya, S., S. E. Hall, et al. (2008). "Agonist-induced conformational changes in bovine rhodopsin: insight into<br />

activation <strong>of</strong> G-protein-coupled receptors." J Mol Biol 382(2): 539-55.<br />

Birnbaumer, L. and M. Birnbaumer (1995). "Signal transduction by G proteins: 1994 edition." J Recept Signal<br />

Transduct Res 15(1-4): 213-52.<br />

Bodor, E. T., G. L. Waldo, et al. (2003). "Purification and functional reconstitution <strong>of</strong> <strong>the</strong> human P2Y12 receptor." Mol<br />

Pharmacol 64(5): 1210-6.<br />

Bogdanov, M., M. Umeda, et al. (1999). "Phospholipid-assisted refolding <strong>of</strong> an integral membrane protein. Minimum<br />

structural features for phosphatidylethanolamine to act as a molecular chaperone." J Biol Chem 274(18):<br />

12339-45.<br />

Bond, R. A., P. Leff, et al. (1995). "Physiological effects <strong>of</strong> inverse agonists in transgenic mice with myocardial<br />

overexpression <strong>of</strong> <strong>the</strong> beta 2-adrenoceptor." Nature 374(6519): 272-6.<br />

Booth, P. J., S. L. Flitsch, et al. (1995). "Intermediates in <strong>the</strong> folding <strong>of</strong> <strong>the</strong> membrane protein bacteriorhodopsin." Nat<br />

Struct Biol 2(2): 139-43.<br />

Borch, J. and T. Hamann (2009). "The nanodisc: a novel tool for membrane protein studies." Biol Chem.<br />

Bosserh<strong>of</strong>f, A., J. Wallach, et al. (1989). "Micropreparative separation <strong>of</strong> peptides derived from sodium dodecyl<br />

sulphate-solubilized proteins." J Chromatogr 473(1): 71-7.<br />

Bourne, H. R. and E. C. Meng (2000). "Structure. Rhodopsin sees <strong>the</strong> light." Science 289(5480): 733-4.<br />

133 | P a g e


References<br />

Bradford, M. M. (1976). "A rapid and sensitive method for <strong>the</strong> quantitation <strong>of</strong> microgram quantities <strong>of</strong> protein utilizing<br />

<strong>the</strong> principle <strong>of</strong> protein-dye binding." Anal Biochem 72: 248-54.<br />

Bradford, M. M. (1976). "A rapid and sensitive method for <strong>the</strong> quantitation <strong>of</strong> microgram quantities <strong>of</strong> protein utilizing<br />

<strong>the</strong> principle <strong>of</strong> protein-dye binding." Anal. Biochem. 72: 248-54.<br />

Bruysters, M., A. Jongejan, et al. (2005). "Pharmacological differences between human and guinea pig histamine H1<br />

receptors: Asn84 (2.61) as key residue within an additional binding pocket in <strong>the</strong> H1 receptor." Mol<br />

Pharmacol 67(4): 1045-52.<br />

Bruysters, M., H. H. Pertz, et al. (2004). "Mutational analysis <strong>of</strong> <strong>the</strong> histamine H1-receptor binding pocket <strong>of</strong><br />

histaprodifens." Eur J Pharmacol 487(1-3): 55-63.<br />

Bullock, A. N., J. Debreczeni, et al. (2005). "Structure and substrate specificity <strong>of</strong> <strong>the</strong> Pim-1 kinase." J Biol Chem<br />

280(50): 41675-82.<br />

Busuttil, B. E., K. L. Turney, et al. (2001). "The expression <strong>of</strong> soluble, full-length, recombinant human TSH receptor in a<br />

prokaryotic system." Protein Expr Purif 23(3): 369-73.<br />

Bywater, R. P. (2005). "Location and nature <strong>of</strong> <strong>the</strong> residues important for ligand recognition in G-protein coupled<br />

receptors." J Mol Recognit 18(1): 60-72.<br />

Canutescu, A. A., A. A. Shelenkov, et al. (2003). "A graph-<strong>the</strong>ory algorithm for rapid protein side-chain prediction."<br />

Protein Sci. 12(9): 2001-14.<br />

Cappuccio, J. A., A. K. Hinz, et al. (2009). "Cell-free expression for nanolipoprotein particles: building a high-throughput<br />

membrane protein solubility platform." Methods Mol Biol 498: 273-96.<br />

Chattopadhyay, A. and E. London (1984). "Fluorimetric determination <strong>of</strong> critical micelle concentration avoiding<br />

interference from detergent charge." Anal Biochem 139(2): 408-12.<br />

Chelikani, P., P. J. Reeves, et al. (2006). "The syn<strong>the</strong>sis and high-level expression <strong>of</strong> a beta2-adrenergic receptor gene in<br />

a tetracycline-inducible stable mammalian cell line." Protein Sci 15(6): 1433-40.<br />

Chen, E. I., D. Cociorva, et al. (2007). "Optimization <strong>of</strong> mass spectrometry-compatible surfactants for shotgun<br />

proteomics." J Proteome Res 6(7): 2529-38.<br />

Chen, S., F. Lin, et al. (2002). "Phe(303) in TMVI <strong>of</strong> <strong>the</strong> alpha(1B)-adrenergic receptor is a key residue coupling TM<br />

helical movements to G-protein activation." Biochemistry 41(2): 588-96.<br />

Cherezov, V., D. M. Rosenbaum, et al. (2007). "High-resolution crystal structure <strong>of</strong> an engineered human beta2adrenergic<br />

G protein-coupled receptor." Science 318(5854): 1258-65.<br />

Christ<strong>of</strong>fers, K. H., H. Li, et al. (2005). "Purification and mass spectrometric analysis <strong>of</strong> <strong>the</strong> delta opioid receptor." Brain<br />

Res Mol Brain Res 136(1-2): 54-64.<br />

Christ<strong>of</strong>fers, K. H., H. Li, et al. (2003). "Purification and mass spectrometric analysis <strong>of</strong> <strong>the</strong> mu opioid receptor." Brain<br />

Res Mol Brain Res 118(1-2): 119-31.<br />

Cianchi, F., C. Cortesini, et al. (2005). "The role <strong>of</strong> cyclooxygenase-2 in mediating <strong>the</strong> effects <strong>of</strong> histamine on cell<br />

proliferation and vascular endo<strong>the</strong>lial growth factor production in colorectal cancer." Clin Cancer Res 11(19<br />

Pt 1): 6807-15.<br />

Condreay, J. P., S. M. Wi<strong>the</strong>rspoon, et al. (1999). "Transient and stable gene expression in mammalian cells transduced<br />

with a recombinant baculovirus vector." Proc Natl Acad Sci U S A 96(1): 127-32.<br />

Cook, D. A. (1971). "Blockade by phenoxybenzamine <strong>of</strong> <strong>the</strong> contractor response produced by agonists in <strong>the</strong> isolated<br />

ileum <strong>of</strong> <strong>the</strong> guinea pig." Br J Pharmacol 43: 197-209.<br />

Cook, D. A., Young, M.S., Ramij, K., Vollrath, B. (1988). "The action <strong>of</strong> agonists and antagonists at <strong>the</strong> histamine H1<br />

receptor and receptor protection studies in guinea pig ileum." Can J Physiol Pharmacol 66: 618-623.<br />

Cormack, B. (2001). "Directed mutagenesis using <strong>the</strong> polymerase chain reaction." Curr. Protoc. Neurosci. Chapter 4:<br />

Unit 4 11.<br />

Costello, C. A., N. L. Kelleher, et al. (1996). "Mechanistic studies on thiaminase I. Overexpression and identification <strong>of</strong><br />

<strong>the</strong> active site nucleophile." J Biol Chem 271(7): 3445-52.<br />

Crocker, E., M. Eilers, et al. (2006). "Location <strong>of</strong> Trp265 in metarhodopsin II: implications for <strong>the</strong> activation<br />

mechanism <strong>of</strong> <strong>the</strong> visual receptor rhodopsin." J. Mol. Biol. 357(1): 163-72.<br />

Cummings, R. D. and S. Kornfeld (1982). "Fractionation <strong>of</strong> asparagine-linked oligosaccharides by serial lectin-Agarose<br />

affinity chromatography. A rapid, sensitive, and specific technique." J Biol Chem 257(19): 11235-40.<br />

Daaka, Y., L. M. Luttrell, et al. (1997). "Switching <strong>of</strong> <strong>the</strong> coupling <strong>of</strong> <strong>the</strong> beta2-adrenergic receptor to different G<br />

proteins by protein kinase A." Nature 390(6655): 88-91.<br />

Daulat, A. M., P. Maurice, et al. (2007). "Purification and identification <strong>of</strong> G protein-coupled receptor protein complexes<br />

under native conditions." Mol Cell Proteomics 6(5): 835-44.<br />

134 | P a g e


References<br />

Davis, D., X. Liu, et al. (1995). "Identification <strong>of</strong> <strong>the</strong> sites <strong>of</strong> N-linked glycosylation on <strong>the</strong> follicle-stimulating hormone<br />

(FSH) receptor and assessment <strong>of</strong> <strong>the</strong>ir role in FSH receptor function." Mol Endocrinol 9(2): 159-70.<br />

de Esch, I. J., R. L. Thurmond, et al. (2005). "The histamine H(4) receptor as a new <strong>the</strong>rapeutic target for<br />

inflammation." Trends Pharmacol Sci 26(9): 462-9.<br />

de Jong, L. A., S. Grunewald, et al. (2004). "Purification and characterization <strong>of</strong> <strong>the</strong> recombinant human dopamine D2S<br />

receptor from Pichia pastoris." Protein Expr Purif 33(2): 176-84.<br />

de Ruyter, P. G., O. P. Kuipers, et al. (1996). "Controlled gene expression systems for Lactococcus lactis with <strong>the</strong> foodgrade<br />

inducer nisin." Appl Environ Microbiol 62(10): 3662-7.<br />

Defea, K. (2008). "Beta-arrestins and heterotrimeric G-proteins: collaborators and competitors in signal<br />

transduction." Br J Pharmacol 153 Suppl 1: S298-309.<br />

Derouazi, M., P. Girard, et al. (2004). "Serum-free large-scale transient transfection <strong>of</strong> CHO cells." Biotechnol Bioeng<br />

87(4): 537-45.<br />

Deslauriers, B., C. Ponce, et al. (1999). "N-glycosylation requirements for <strong>the</strong> AT1a angiotensin II receptor delivery to<br />

<strong>the</strong> plasma membrane." Biochem J 339 ( Pt 2): 397-405.<br />

Deupi, X., N. Dolker, et al. (2007). "Structural Models <strong>of</strong> Class A G protein-coupled receptors as a Tool for Drug Design:<br />

Insights on Transmembrane Bundle Plasticity." Curr. Top. Med. Chem. 7(10): 999-1006.<br />

Deupi, X., M. Olivella, et al. (2004). "Ser and Thr residues modulate <strong>the</strong> conformation <strong>of</strong> pro-kinked transmembrane<br />

alpha-helices." Biophys. J. 86(1): 105-15.<br />

Dolder, M., A. Engel, et al. (1996). "The micelle to vesicle transition <strong>of</strong> lipids and detergents in <strong>the</strong> presence <strong>of</strong> a<br />

membrane protein: towards a rationale for 2D crystallization." FEBS Lett 382(1-2): 203-8.<br />

Donaldson, J. and S. J. Hill (1986). "Selective enhancement <strong>of</strong> histamine H1-receptor responses in guinea-pig ileal<br />

smooth muscle by 1,4-dithiothreitol." Br J Pharmacol 87(1): 191-9.<br />

Dong, C., C. M. Filipeanu, et al. (2007). "Regulation <strong>of</strong> G protein-coupled receptor export trafficking." Biochim Biophys<br />

Acta 1768(4): 853-70.<br />

Donnelly, D., J. B. Findlay, et al. (1994). "The evolution and structure <strong>of</strong> aminergic G protein-coupled receptors."<br />

Receptors Channels 2(1): 61-78.<br />

Doucette, A., D. Craft, et al. (2000). "Protein concentration and enzyme digestion on microbeads for MALDI-TOF<br />

peptides mass mapping <strong>of</strong> proteins from dilute solutions." Anal Chem 72(14): 3355-62.<br />

Dryja, T. P., E. L. Berson, et al. (1993). "Heterozygous missense mutation in <strong>the</strong> rhodopsin gene as a cause <strong>of</strong> congenital<br />

stationary night blindness." Nat Genet 4(3): 280-3.<br />

Duffin, K. L., G. W. Lange, et al. (1993). "Identification and oligosaccharide structure analysis <strong>of</strong> rhodopsin glyc<strong>of</strong>orms<br />

containing galactose and sialic acid." Glycobiology 3(4): 365-80.<br />

Dunham, T. D. and D. L. Farrens (1999). "Conformational changes in rhodopsin. Movement <strong>of</strong> helix f detected by sitespecific<br />

chemical labeling and fluorescence spectroscopy." J Biol Chem 274(3): 1683-90.<br />

Eckart, M. R. and C. M. Bussineau (1996). "Quality and au<strong>the</strong>nticity <strong>of</strong> heterologous proteins syn<strong>the</strong>sized in yeast."<br />

Curr Opin Biotechnol 7(5): 525-30.<br />

Edwards, P. C., J. Li, et al. (2004). "Crystals <strong>of</strong> native and modified bovine rhodopsins and <strong>the</strong>ir heavy atom<br />

derivatives." J Mol Biol 343(5): 1439-50.<br />

Ekstrom, S., J. Malmstrom, et al. (2002). "On-chip microextraction for proteomic sample preparation <strong>of</strong> in-gel digests."<br />

Proteomics 2(4): 413-21.<br />

Elbaz, Y., S. Steiner-Mordoch, et al. (2004). "In vitro syn<strong>the</strong>sis <strong>of</strong> fully functional EmrE, a multidrug transporter, and<br />

study <strong>of</strong> its oligomeric state." Proc Natl Acad Sci U S A 101(6): 1519-24.<br />

Elz, S., K. Kramer, et al. (2000). "Histaprodifens: syn<strong>the</strong>sis, pharmacological in vitro evaluation, and molecular<br />

modeling <strong>of</strong> a new class <strong>of</strong> highly active and selective histamine H(1)-receptor agonists." J Med Chem 43(6):<br />

1071-84.<br />

Engel, A., A. Hoenger, et al. (1992). "Assembly <strong>of</strong> 2-D membrane protein crystals: dynamics, crystal order, and fidelity<br />

<strong>of</strong> structure analysis by electron microscopy." J Struct Biol 109(3): 219-34.<br />

Eroglu, C., P. Cronet, et al. (2002). "Functional reconstitution <strong>of</strong> purified metabotropic glutamate receptor expressed in<br />

<strong>the</strong> fly eye." EMBO Rep 3(5): 491-6.<br />

Espinoza-Fonseca, L. M., Trujillo-Ferrara, J.G., (2006). "The existence <strong>of</strong> a second allosteric site on <strong>the</strong> M1 muscarinic<br />

acetylcholine receptor and its implications for drug design." Bioorg Med Chem Lett 16: 1217-1220.<br />

Feng, W., J. Cai, et al. (2002). "Expression <strong>of</strong> CB2 cannabinoid receptor in Pichia pastoris." Protein Expr Purif 26(3):<br />

496-505.<br />

Ficca, A. G., L. Testa, et al. (1995). "The human beta 2-adrenergic receptor expressed in Schizosaccharomyces pombe<br />

retains its pharmacological properties." FEBS Lett 377(2): 140-4.<br />

135 | P a g e


References<br />

Filppula, S., S. Yaddanapudi, et al. (2004). "Purification and mass spectroscopic analysis <strong>of</strong> human CB2 cannabinoid<br />

receptor expressed in <strong>the</strong> baculovirus system." J Pept Res 64(6): 225-36.<br />

Fitzsimons, C. P., F. Monczor, et al. (2004). "Mepyramine, a histamine H1 receptor inverse agonist, binds preferentially<br />

to a G protein-coupled form <strong>of</strong> <strong>the</strong> receptor and sequesters G protein." J Biol Chem 279(33): 34431-9.<br />

Francken, B., W. Luyten, et al. (1997). Molecular Mechanisms <strong>of</strong> Signalling and Membrane Transport. Berlin, Springer-<br />

Verlag.<br />

Frang, H., V. Cockcr<strong>of</strong>t, et al. (2001). "Phenoxybenzamine binding reveals <strong>the</strong> helical orientation <strong>of</strong> <strong>the</strong> third<br />

transmembrane domain <strong>of</strong> adrenergic receptors." J Biol Chem 276(33): 31279-84.<br />

Fraser, M. J. (1992). "The baculovirus-infected insect cell as a eukaryotic gene expression system." Curr Top Microbiol<br />

Immunol 158: 131-72.<br />

Fukui, H., N. P. Wang, et al. (1988). "Solubilization, characterization and partial purification <strong>of</strong> [3H]mepyraminebinding<br />

protein, a possible histamine H1 receptor, from rat liver membrane." Jpn J Pharmacol 46(2): 127-39.<br />

Fung-Leung, W. P., R. L. Thurmond, et al. (2004). "Histamine H4 receptor antagonists: <strong>the</strong> new antihistamines?" Curr<br />

Opin Investig Drugs 5(11): 1174-83.<br />

Furukawa, H. and T. Haga (2000). "Expression <strong>of</strong> functional M2 muscarinic acetylcholine receptor in Escherichia coli."<br />

J Biochem 127(1): 151-61.<br />

Gan, L., J. M. Alexander, et al. (2006). "Large-scale purification and characterization <strong>of</strong> human parathyroid hormone-1<br />

receptor stably expressed in HEK293S GnTI- cells." Protein Expr Purif 47(1): 296-302.<br />

Geisse, S., M. Jordan, et al. (2005). "Large-scale transient expression <strong>of</strong> <strong>the</strong>rapeutic proteins in mammalian cells."<br />

Methods Mol Biol 308: 87-98.<br />

Ge<strong>the</strong>r, U., S. Lin, et al. (1997). "Agonists induce conformational changes in transmembrane domains III and VI <strong>of</strong> <strong>the</strong><br />

b2 adrenergic receptor." EMBO J. 16(22): 6737-47.<br />

Ghanouni, P., H. Schambye, et al. (2000). "The effect <strong>of</strong> pH on beta(2) adrenoceptor function. Evidence for protonationdependent<br />

activation." J Biol Chem 275(5): 3121-7.<br />

Giga-Hama, Y. and H. Kumagai (1998). "[Foreign gene expression in fission yeast S. pombe]." Seikagaku 70(4): 300-4.<br />

Gillard, M., C. Van Der Perren, et al. (2002). "Binding characteristics <strong>of</strong> cetirizine and levocetirizine to human H(1)<br />

histamine receptors: contribution <strong>of</strong> Lys(191) and Thr(194)." Mol Pharmacol 61(2): 391-9.<br />

Gimpl, G. and F. Fahrenholz (2002). "Cholesterol as stabilizer <strong>of</strong> <strong>the</strong> oxytocin receptor." Biochim Biophys Acta<br />

1564(2): 384-92.<br />

Gimpl, G., U. Klein, et al. (1995). "Expression <strong>of</strong> <strong>the</strong> human oxytocin receptor in baculovirus-infected insect cells: highaffinity<br />

binding is induced by a cholesterol-cyclodextrin complex." Biochemistry 34(42): 13794-801.<br />

Girault, S., S. Sagan, et al. (1996). "The use <strong>of</strong> photolabelled peptides to localize <strong>the</strong> substance-P-binding site in <strong>the</strong><br />

human neurokinin-1 tachykinin receptor." Eur J Biochem 240(1): 215-22.<br />

Goldman, L. A., E. C. Cutrone, et al. (1996). "Modifications <strong>of</strong> vectors pEF-BOS, pcDNA1 and pcDNA3 result in improved<br />

convenience and expression." Biotechniques 21(6): 1013-5.<br />

Govaerts, C., A. Lefort, et al. (2001). "A conserved Asn in transmembrane helix 7 is an on/<strong>of</strong>f switch in <strong>the</strong> activation <strong>of</strong><br />

<strong>the</strong> thyrotropin receptor." J Biol Chem 276(25): 22991-9.<br />

Graslund, S., P. Nordlund, et al. (2008). "Protein production and purification." Nat Methods 5(2): 135-46.<br />

Grisshammer, R. (2006). "Understanding recombinant expression <strong>of</strong> membrane proteins." Curr Opin Biotechnol<br />

17(4): 337-40.<br />

Grisshammer, R., R. Duckworth, et al. (1993). "Expression <strong>of</strong> a rat neurotensin receptor in Escherichia coli." Biochem J<br />

295 ( Pt 2): 571-6.<br />

Grisshammer, R., J. Little, et al. (1994). "Expression <strong>of</strong> rat NK-2 (neurokinin A) receptor in E. coli." Receptors Channels<br />

2(4): 295-302.<br />

Grisshammer, R. and C. G. Tate (1995). "Overexpression <strong>of</strong> integral membrane proteins for structural studies." Q Rev<br />

Biophys 28(3): 315-422.<br />

Grisshammer, R. and J. Tucker (1997). "Quantitative evaluation <strong>of</strong> neurotensin receptor purification by immobilized<br />

metal affinity chromatography." Protein Expr Purif 11(1): 53-60.<br />

Grisshammer, R., J. F. White, et al. (2005). "Large-scale expression and purification <strong>of</strong> a G-protein-coupled receptor for<br />

structure determination -- an overview." J Struct Funct Genomics 6(2-3): 159-63.<br />

Gwyn, W. G. (1994). Membrane protein expression systems. London, Portland Press.<br />

Haendler, B., U. Hechler, et al. (1993). "Expression <strong>of</strong> human endo<strong>the</strong>lin receptor ETB by Escherichia coli<br />

transformants." Biochem Biophys Res Commun 191(2): 633-8.<br />

Hale, C. C., R. S. Slaughter, et al. (1984). "Identification and partial purification <strong>of</strong> <strong>the</strong> cardiac sodium-calcium exchange<br />

protein." Proc Natl Acad Sci U S A 81(21): 6569-73.<br />

136 | P a g e


References<br />

Hamilton, S. R., R. C. Davidson, et al. (2006). "Humanization <strong>of</strong> yeast to produce complex terminally sialylated<br />

glycoproteins." Science 313(5792): 1441-3.<br />

Hampe, W., R. H. Voss, et al. (2000). "Engineering <strong>of</strong> a proteolytically stable human beta 2-adrenergic<br />

receptor/maltose-binding protein fusion and production <strong>of</strong> <strong>the</strong> chimeric protein in Escherichia coli and<br />

baculovirus-infected insect cells." J Biotechnol 77(2-3): 219-34.<br />

Han, J. and K. L. Schey (2004). "Proteolysis and mass spectrometric analysis <strong>of</strong> an integral membrane: aquaporin 0." J<br />

Proteome Res 3(4): 807-12.<br />

Han, M., S. W. Lin, et al. (1996). "Functional interaction <strong>of</strong> transmembrane helices 3 and 6 in rhodopsin. Replacement<br />

<strong>of</strong> phenylalanine 261 by alanine causes reversion <strong>of</strong> phenotype <strong>of</strong> a glycine 121 replacement mutant." J Biol<br />

Chem 271(50): 32337-42.<br />

Hanson, M. A., V. Cherezov, et al. (2008). "A specific cholesterol binding site is established by <strong>the</strong> 2.8 A structure <strong>of</strong> <strong>the</strong><br />

human beta2-adrenergic receptor." Structure 16(6): 897-905.<br />

Hanson, M. A. and R. C. Stevens (2009). "Discovery <strong>of</strong> new GPCR biology: one receptor structure at a time." Structure<br />

17(1): 8-14.<br />

Harding, P. J., H. Attrill, et al. (2007). "Neurotensin receptor type 1: Escherichia coli expression, purification,<br />

characterization and biophysical study." Biochem Soc Trans 35(Pt 4): 760-3.<br />

Hasegawa, J., H. H. Loh, et al. (1987). "Lipid requirement for mu opioid receptor binding." J Neurochem 49(4): 1007-<br />

12.<br />

Hassaine, G., R. Wagner, et al. (2006). "Semliki Forest virus vectors for overexpression <strong>of</strong> 101 G protein-coupled<br />

receptors in mammalian host cells." Protein Expr Purif 45(2): 343-51.<br />

Hayashi, M. K. and T. Haga (1996). "Purification and functional reconstitution with GTP-binding regulatory proteins <strong>of</strong><br />

hexahistidine-tagged muscarinic acetylcholine receptors (m2 subtype)." J Biochem 120(6): 1232-8.<br />

Ho, J. T., J. F. White, et al. (2008). "Analysis <strong>of</strong> a G protein-coupled receptor for neurotensin by liquid chromatographyelectrospray<br />

ionization-mass spectrometry." Anal Biochem 376(1): 13-24.<br />

H<strong>of</strong>mann, C., V. Sandig, et al. (1995). "Efficient gene transfer into human hepatocytes by baculovirus vectors." Proc<br />

Natl Acad Sci U S A 92(22): 10099-103.<br />

Horn, F., E. Bettler, et al. (2003). "GPCRDB information system for G protein-coupled receptors." Nucleic Acids Res<br />

31(1): 294-7.<br />

Huang, P., J. Li, et al. (2001). "Functional role <strong>of</strong> a conserved motif in TM6 <strong>of</strong> <strong>the</strong> rat mu opioid receptor: constitutively<br />

active and inactive receptors result from substitutions <strong>of</strong> Thr6.34(279) with Lys and Asp." Biochemistry<br />

40(45): 13501-9.<br />

Huang, P., I. Visiers, et al. (2002). "The local environment at <strong>the</strong> cytoplasmic end <strong>of</strong> TM6 <strong>of</strong> <strong>the</strong> mu opioid receptor<br />

differs from those <strong>of</strong> rhodopsin and monoamine receptors: introduction <strong>of</strong> an ionic lock between <strong>the</strong><br />

cytoplasmic ends <strong>of</strong> helices 3 and 6 by a L6.30(275)E mutation inactivates <strong>the</strong> mu opioid receptor and<br />

reduces <strong>the</strong> constitutive activity <strong>of</strong> its T6.34(279)K mutant." Biochemistry 41(40): 11972-80.<br />

Hulme, E. C. and C. A. Curtis (1998). "Purification <strong>of</strong> recombinant M1 muscarinic acetylcholine receptor." Biochem Soc<br />

Trans 26(4): S361.<br />

Ikawa, Y., M. Suzuki, et al. (2005). "Histamine H4 receptor expression in human synovial cells obtained from patients<br />

suffering from rheumatoid arthritis." Biol Pharm Bull 28(10): 2016-8.<br />

Inoue, I., K. Yanai, et al. (1996). "Impaired locomotor activity and exploratory behavior in mice lacking histamine H1<br />

receptors." Proc Natl Acad Sci U S A 93(23): 13316-20.<br />

Ishihara, G., M. Goto, et al. (2005). "Expression <strong>of</strong> G protein coupled receptors in a cell-free translational system using<br />

detergents and thioredoxin-fusion vectors." Protein Expr Purif 41(1): 27-37.<br />

Ivanova, L., J. Brandli, et al. (2005). "Hybrid Sindbis/Epstein-Barr virus episomal expression vector for inducible<br />

production <strong>of</strong> proteins." Biotechniques 39(2): 209-12.<br />

Jaakola, V. P., M. T. Griffith, et al. (2008). "The 2.6 angstrom crystal structure <strong>of</strong> a human A2A adenosine receptor<br />

bound to an antagonist." Science 322(5905): 1211-7.<br />

Jacoby, E., R. Bouhelal, et al. (2006). "The 7 TM G-protein-coupled receptor target family." ChemMedChem 1(8): 761-<br />

82.<br />

Janssen, J. J., P. H. Bovee-Geurts, et al. (1995). "Histidine tagging both allows convenient single-step purification <strong>of</strong><br />

bovine rhodopsin and exerts ionic strength-dependent effects on its photochemistry." J Biol Chem 270(19):<br />

11222-9.<br />

Jap, B. K., P. J. Walian, et al. (1991). "Structural architecture <strong>of</strong> an outer membrane channel as determined by electron<br />

crystallography." Nature 350(6314): 167-70.<br />

137 | P a g e


References<br />

Jayadev, S., R. D. Smith, et al. (1999). "N-linked glycosylation is required for optimal AT1a angiotensin receptor<br />

expression in COS-7 cells." Endocrinology 140(5): 2010-7.<br />

Jenkinson, D. (1996). Textbook <strong>of</strong> Receptor pharmacology. Boca Raton, FL, CRC Press, Inc.<br />

Jockers, R., M. E. Linder, et al. (1994). "Species difference in <strong>the</strong> G protein selectivity <strong>of</strong> <strong>the</strong> human and bovine A1adenosine<br />

receptor." J Biol Chem 269(51): 32077-84.<br />

Jongejan, A., M. Bruysters, et al. (2005). "Linking agonist binding to histamine H1 receptor activation." Nat Chem Biol<br />

1(2): 98-103.<br />

Jongejan, A. and R. Leurs (2005). "Delineation <strong>of</strong> receptor-ligand interactions at <strong>the</strong> human histamine H1 receptor by a<br />

combined approach <strong>of</strong> site-directed mutagenesis and computational techniques - or - how to bind <strong>the</strong> H1<br />

receptor." Arch Pharm (Weinheim) 338(5-6): 248-59.<br />

Kamonchanok, S., C. I. Balog, et al. (2008). "GPCR proteomics: mass spectrometric and functional analysis <strong>of</strong> histamine<br />

H1 receptor after baculovirus-driven and in vitro cell free expression." J Proteome Res 7(2): 621-9.<br />

Katzen, F., T. C. Peterson, et al. (2009). "Membrane protein expression: no cells required." Trends Biotechnol 27(8):<br />

455-60.<br />

Kaushal, S., K. D. Ridge, et al. (1994). "Structure and function in rhodopsin: <strong>the</strong> role <strong>of</strong> asparagine-linked glycosylation."<br />

Proc Natl Acad Sci U S A 91(9): 4024-8.<br />

Kawarasaki, Y., T. Kawai, et al. (1995). "A long-lived batch reaction system <strong>of</strong> cell-free protein syn<strong>the</strong>sis." Anal<br />

Biochem 226(2): 320-4.<br />

Kelley, G. G., K. A. Kaproth-Joslin, et al. (2006). "G-protein-coupled receptor agonists activate endogenous<br />

phospholipase Cepsilon and phospholipase Cbeta3 in a temporally distinct manner." J Biol Chem 281(5):<br />

2639-48.<br />

Kenakin, T. (2007). "Collateral efficacy in drug discovery: taking advantage <strong>of</strong> <strong>the</strong> good (allosteric) nature <strong>of</strong> 7TM<br />

receptors." Trends Pharmacol. Sci. 28(8): 407-15.<br />

Kiefer, H., J. Krieger, et al. (1996). "Expression <strong>of</strong> an olfactory receptor in Escherichia coli: purification, reconstitution,<br />

and ligand binding." Biochemistry 35(50): 16077-84.<br />

Kigawa, T., Y. Muto, et al. (1995). "Cell-free syn<strong>the</strong>sis and amino acid-selective stable isotope labeling <strong>of</strong> proteins for<br />

NMR analysis." J Biomol NMR 6(2): 129-34.<br />

Kigawa, T., T. Yabuki, et al. (1999). "Cell-free production and stable-isotope labeling <strong>of</strong> milligram quantities <strong>of</strong><br />

proteins." FEBS Lett 442(1): 15-9.<br />

Kigawa, T. and S. Yokoyama (1991). "A continuous cell-free protein syn<strong>the</strong>sis system for coupled transcriptiontranslation."<br />

J Biochem 110(2): 166-8.<br />

Kim, D. M., T. Kigawa, et al. (1996). "A highly efficient cell-free protein syn<strong>the</strong>sis system from Escherichia coli." Eur J<br />

Biochem 239(3): 881-6.<br />

Kim, J. M., C. Altenbach, et al. (1997). "Structure and function in rhodopsin: rhodopsin mutants with a neutral amino<br />

acid at E134 have a partially activated conformation in <strong>the</strong> dark state." Proc Natl Acad Sci U S A 94(26):<br />

14273-8.<br />

Kim, T. K., R. Zhang, et al. (2005). "Expression and characterization <strong>of</strong> human CB1 cannabinoid receptor in<br />

methylotrophic yeast Pichia pastoris." Protein Expr Purif 40(1): 60-70.<br />

Kim, T. W., I. S. Oh, et al. (2006). "Cell-free syn<strong>the</strong>sis and in situ isolation <strong>of</strong> recombinant proteins." Protein Expr Purif<br />

45(2): 249-54.<br />

Klaassen, C. H., P. H. Bovee-Geurts, et al. (1999). "Large-scale production and purification <strong>of</strong> functional recombinant<br />

bovine rhodopsin with <strong>the</strong> use <strong>of</strong> <strong>the</strong> baculovirus expression system." Biochem J 342 ( Pt 2): 293-300.<br />

Klammt, C., F. Lohr, et al. (2004). "High level cell-free expression and specific labeling <strong>of</strong> integral membrane proteins."<br />

Eur J Biochem 271(3): 568-80.<br />

Klammt, C., D. Schwarz, et al. (2007). "Cell-free production <strong>of</strong> G protein-coupled receptors for functional and structural<br />

studies." J Struct Biol.<br />

Klammt, C., D. Schwarz, et al. (2005). "Evaluation <strong>of</strong> detergents for <strong>the</strong> soluble expression <strong>of</strong> alpha-helical and betabarrel-type<br />

integral membrane proteins by a preparative scale individual cell-free expression system." Febs J<br />

272(23): 6024-38.<br />

Klammt, C., D. Schwarz, et al. (2006). "Cell-free expression as an emerging technique for <strong>the</strong> large scale production <strong>of</strong><br />

integral membrane protein." Febs J 273(18): 4141-53.<br />

Kleymann, G., F. Boege, et al. (1993). "Human beta 2-adrenergic receptor produced in stably transformed insect cells<br />

is functionally coupled via endogenous GTP-binding protein to adenylyl cyclase." Eur J Biochem 213(2): 797-<br />

804.<br />

138 | P a g e


References<br />

Knight, P. J., T. A. Pfeifer, et al. (2003). "A functional assay for G-protein-coupled receptors using stably transformed<br />

insect tissue culture cell lines." Anal Biochem 320(1): 88-103.<br />

Kobilka, B. (2004). "Agonist binding: a multistep process." Mol. Pharmacol. 65(5): 1060-2.<br />

Kobilka, B. and G. F. Schertler (2008). "New G-protein-coupled receptor crystal structures: insights and limitations."<br />

Trends Pharmacol Sci 29(2): 79-83.<br />

Kobilka, B. K. (1995). "Amino and carboxyl terminal modifications to facilitate <strong>the</strong> production and purification <strong>of</strong> a G<br />

protein-coupled receptor." Anal Biochem 231(1): 269-71.<br />

Kobilka, B. K. and X. Deupi (2007). "Conformational complexity <strong>of</strong> G-protein-coupled receptors." Trends Pharmacol<br />

Sci 28(8): 397-406.<br />

Kohno, M., N. Fukushima, et al. (2000). "G(i1) and G(oA) differentially determine kinetic efficacies <strong>of</strong> agonists for<br />

kappa-opioid receptor." FEBS Lett 473(1): 101-5.<br />

Kohno, T., A. Wada, et al. (2002). "N-glycans <strong>of</strong> sphingosine 1-phosphate receptor Edg-1 regulate ligand-induced<br />

receptor internalization." Faseb J 16(9): 983-92.<br />

Kraft, P., J. Mills, et al. (2001). "Mass spectrometric analysis <strong>of</strong> cyanogen bromide fragments <strong>of</strong> integral membrane<br />

proteins at <strong>the</strong> picomole level: application to rhodopsin." Anal Biochem 292(1): 76-86.<br />

Krepkiy, D., K. Wong, et al. (2006). "Bacterial expression <strong>of</strong> functional, biotinylated peripheral cannabinoid receptor<br />

CB2." Protein Expr Purif 49(1): 60-70.<br />

Kristiansen, K. (2004). "Molecular mechanisms <strong>of</strong> ligand binding, signaling, and regulation within <strong>the</strong> superfamily <strong>of</strong><br />

G-protein-coupled receptors: molecular modeling and mutagenesis approaches to receptor structure and<br />

function." Pharmacol. Ther. 103(1): 21-80.<br />

Kuhlbrandt, W., D. N. Wang, et al. (1994). "Atomic model <strong>of</strong> plant light-harvesting complex by electron<br />

crystallography." Nature 367(6464): 614-21.<br />

Kuhn, B. and T. Gudermann (1999). "The luteinizing hormone receptor activates phospholipase C via preferential<br />

coupling to Gi2." Biochemistry 38(38): 12490-8.<br />

Kunji, E. R., D. J. Slotboom, et al. (2003). "Lactococcus lactis as host for overproduction <strong>of</strong> functional membrane<br />

proteins." Biochim Biophys Acta 1610(1): 97-108.<br />

Kusui, T., R. V. Benya, et al. (1994). "Glycosylation <strong>of</strong> bombesin receptors: characterization, effect on binding, and Gprotein<br />

coupling." Biochemistry 33(44): 12968-80.<br />

Kwatra, M. M., J. Schreurs, et al. (1995). "Immunoaffinity purification <strong>of</strong> epitope-tagged human beta 2-adrenergic<br />

receptor to homogeneity." Protein Expr Purif 6(6): 717-21.<br />

Kwatra, M. M., D. A. Schwinn, et al. (1993). "The substance P receptor, which couples to Gq/11, is a substrate <strong>of</strong> betaadrenergic<br />

receptor kinase 1 and 2." J Biol Chem 268(13): 9161-4.<br />

Lagerstrom, M. C. and H. B. Schioth (2008). "Structural diversity <strong>of</strong> G protein-coupled receptors and significance for<br />

drug discovery." Nat Rev Drug Discov 7(4): 339-57.<br />

Lambert, O., D. Levy, et al. (1998). "A new "gel-like" phase in dodecyl maltoside-lipid mixtures: implications in<br />

solubilization and reconstitution studies." Biophys J 74(2 Pt 1): 918-30.<br />

Lau, F. W. and J. U. Bowie (1997). "A method for assessing <strong>the</strong> stability <strong>of</strong> a membrane protein." Biochemistry 36(19):<br />

5884-92.<br />

Laue, L., W. Y. Chan, et al. (1995). "Genetic heterogeneity <strong>of</strong> constitutively activating mutations <strong>of</strong> <strong>the</strong> human<br />

luteinizing hormone receptor in familial male-limited precocious puberty." Proc Natl Acad Sci U S A 92(6):<br />

1906-10.<br />

le Maire, M., P. Champeil, et al. (2000). "Interaction <strong>of</strong> membrane proteins and lipids with solubilizing detergents."<br />

Biochim Biophys Acta 1508(1-2): 86-111.<br />

Lefkowitz, R. J. (2007). "Seven transmembrane receptors: something old, something new." Acta Physiol (Oxf) 190(1):<br />

9-19.<br />

Lefkowitz, R. J. and S. K. Shenoy (2005). "Transduction <strong>of</strong> receptor signals by beta-arrestins." Science 308(5721): 512-<br />

7.<br />

Leitz, A. J., T. H. Bayburt, et al. (2006). "Functional reconstitution <strong>of</strong> Beta2-adrenergic receptors utilizing selfassembling<br />

Nanodisc technology." Biotechniques 40(5): 601-2, 604, 606, passim.<br />

Leopoldt, D., C. Harteneck, et al. (1997). "G proteins endogenously expressed in Sf9 cells: interactions with mammalian<br />

histamine receptors." Naunyn Schmiedebergs Arch Pharmacol 356: 216-224.<br />

Leschke, C., S. Elz, et al. (1995). "Syn<strong>the</strong>sis and histamine H1 receptor agonist activity <strong>of</strong> a series <strong>of</strong> 2phenylhistamines,<br />

2-heteroarylhistamines, and analogues." J Med Chem 38(8): 1287-94.<br />

Leurs, R., M. K. Church, et al. (2002). "H1-antihistamines: inverse agonism, anti-inflammatory actions and cardiac<br />

effects." Clin Exp Allergy 32(4): 489-98.<br />

139 | P a g e


References<br />

Leurs, R., M. J. Smit, et al. (1995). "Lysine200 located in <strong>the</strong> fifth transmembrane domain <strong>of</strong> <strong>the</strong> histamine H1 receptor<br />

interacts with histamine but not with all H1 agonists." Biochem Biophys Res Commun 214(1): 110-7.<br />

Leurs, R., M. J. Smit, et al. (1994). "Site-directed mutagenesis <strong>of</strong> <strong>the</strong> histamine H1-receptor reveals a selective<br />

interaction <strong>of</strong> asparagine207 with subclasses <strong>of</strong> H1-receptor agonists." Biochem Biophys Res Commun<br />

201(1): 295-301.<br />

Levy, D., A. Bluzat, et al. (1990). "A systematic study <strong>of</strong> liposome and proteoliposome reconstitution involving Bio-<br />

Bead-mediated Triton X-100 removal." Biochim Biophys Acta 1025(2): 179-90.<br />

Li, J., P. C. Edwards, et al. (2004). "Structure <strong>of</strong> bovine rhodopsin in a trigonal crystal form." J. Mol. Biol. 343(5): 1409-<br />

38.<br />

Li, N., D. Salom, et al. (2007). "Heterologous expression <strong>of</strong> <strong>the</strong> adenosine A1 receptor in transgenic mouse retina."<br />

Biochemistry 46(28): 8350-9.<br />

Liitti, S., M. T. Matikainen, et al. (2001). "Immunoaffinity purification and reconstitution <strong>of</strong> human alpha(2)-adrenergic<br />

receptor subtype C2 into phospholipid vesicles." Protein Expr Purif 22(1): 1-10.<br />

Lipscomb, M. L., M. C. Mowry, et al. (2004). "Production <strong>of</strong> a secreted glycoprotein from an inducible promoter system<br />

in a perfusion bioreactor." Biotechnol Prog 20(5): 1402-7.<br />

Lorch, M., I. Lehner, et al. (2005). "NMR and fluorescence spectroscopy approaches to secondary and primary active<br />

multidrug efflux pumps." Biochem Soc Trans 33(Pt 4): 873-7.<br />

Luecke, H., B. Schobert, et al. (1999). "Structure <strong>of</strong> bacteriorhodopsin at 1.55 A resolution." J Mol Biol 291(4): 899-911.<br />

Lundstrom, K. (2005). "The future <strong>of</strong> G protein-coupled receptors as targets in drug discovery." IDrugs 8(11): 909-13.<br />

Lundstrom, K. (2007). "Structural genomics and drug discovery." J Cell Mol Med 11(2): 224-38.<br />

Lundstrom, K., R. Wagner, et al. (2006). "Structural genomics on membrane proteins: comparison <strong>of</strong> more than 100<br />

GPCRs in 3 expression systems." J Struct Funct Genomics 7(2): 77-91.<br />

Madabushi, S., A. K. Gross, et al. (2004). "Evolutionary trace <strong>of</strong> G protein-coupled receptors reveals clusters <strong>of</strong> residues<br />

that determine global and class-specific functions." J Biol Chem 279(9): 8126-32.<br />

Madin, K., T. Sawasaki, et al. (2000). "A highly efficient and robust cell-free protein syn<strong>the</strong>sis system prepared from<br />

wheat embryos: plants apparently contain a suicide system directed at ribosomes." Proc Natl Acad Sci U S A<br />

97(2): 559-64.<br />

Magnin, T., C. Fiez-Vandal, et al. (2009). "A novel, generic and effective method for <strong>the</strong> rapid purification <strong>of</strong> G proteincoupled<br />

receptors." Protein Expr Purif 64(1): 1-7.<br />

Marheineke, K., S. Grunewald, et al. (1998). "Lipid composition <strong>of</strong> Spodoptera frugiperda (Sf9) and Trichoplusia ni<br />

(Tn) insect cells used for baculovirus infection." FEBS Lett 441(1): 49-52.<br />

Marinissen, M. J. and J. S. Gutkind (2001). "G-protein-coupled receptors and signaling networks: emerging paradigms."<br />

Trends Pharmacol Sci 22(7): 368-76.<br />

Marinissen, M. J., J. M. Servitja, et al. (2003). "Thrombin protease-activated receptor-1 signals through Gq- and G13initiated<br />

MAPK cascades regulating c-Jun expression to induce cell transformation." J Biol Chem 278(47):<br />

46814-25.<br />

Marti-Renom, M. A., A. C. Stuart, et al. (2000). "Comparative protein structure modeling <strong>of</strong> genes and genomes." Annu.<br />

Rev. Biophys. Biomol. Struct. 29: 291-325.<br />

Marullo, S., C. Delavier-Klutchko, et al. (1988). "Human beta 2-adrenergic receptors expressed in Escherichia coli<br />

membranes retain <strong>the</strong>ir pharmacological properties." Proc Natl Acad Sci U S A 85(20): 7551-5.<br />

Massotte, D. (2003). "G protein-coupled receptor overexpression with <strong>the</strong> baculovirus-insect cell system: a tool for<br />

structural and functional studies." Biochim Biophys Acta 1610(1): 77-89.<br />

Mazina, K. E., C. D. Strader, et al. (1996). "Purification and reconstitution <strong>of</strong> a recombinant human neurokinin-1<br />

receptor." J Recept Signal Transduct Res 16(3-4): 191-207.<br />

McCusker, E. C., S. E. Bane, et al. (2007). "Heterologous GPCR expression: a bottleneck to obtaining crystal structures."<br />

Biotechnol Prog 23(3): 540-7.<br />

McGregor, M. J., S. A. Islam, et al. (1987). "Analysis <strong>of</strong> <strong>the</strong> Relationship Between Side-chain Conformation and<br />

Secondary Structure in Globular Proteins." J. Mol. Biol. 198: 295-310.<br />

Menon, S. T., M. Han, et al. (2001). "Rhodopsin: structural basis <strong>of</strong> molecular physiology." Physiol Rev 81(4): 1659-88.<br />

Midgett, C. R. and D. R. Madden (2007). "Breaking <strong>the</strong> bottleneck: Eukaryotic membrane protein expression for highresolution<br />

structural studies." J Struct Biol 160(3): 265-74.<br />

Mills, J. S., H. M. Miettinen, et al. (1998). "Identification <strong>of</strong> a ligand binding site in <strong>the</strong> human neutrophil formyl peptide<br />

receptor using a site-specific fluorescent photoaffinity label and mass spectrometry." J Biol Chem 273(17):<br />

10428-35.<br />

140 | P a g e


References<br />

Mirzadegan, T., G. Benko, et al. (2003). "Sequence analyses <strong>of</strong> G-protein-coupled receptors: similarities to rhodopsin."<br />

Biochemistry 42(10): 2759-67.<br />

Moguilevsky, N., F. Varsalona, et al. (1995). "Pharmacological and functional characterisation <strong>of</strong> <strong>the</strong> wild-type and sitedirected<br />

mutants <strong>of</strong> <strong>the</strong> human H1 histamine receptor stably expressed in CHO cells." J Recept Signal<br />

Transduct Res 15(1-4): 91-102.<br />

Mollaaghababa, R., F. F. Davidson, et al. (1996). "Structure and function in rhodopsin: expression <strong>of</strong> functional<br />

mammalian opsin in Saccharomyces cerevisiae." Proc Natl Acad Sci U S A 93(21): 11482-6.<br />

Monne, M., K. W. Chan, et al. (2005). "Functional expression <strong>of</strong> eukaryotic membrane proteins in Lactococcus lactis."<br />

Protein Sci 14(12): 3048-56.<br />

Mulheron, J. G., S. J. Casanas, et al. (1994). "Human 5-HT1A receptor expressed in insect cells activates endogenous<br />

G(o)-like G protein(s)." J Biol Chem 269(17): 12954-62.<br />

Murakami, M. and T. Kouyama (2008). "Crystal structure <strong>of</strong> squid rhodopsin." Nature 453(7193): 363-7.<br />

Murayama, T., Y. Kajiyama, et al. (1990). "Histamine-stimulated and GTP-binding proteins-mediated phospholipase<br />

A2 activation in rabbit platelets." J Biol Chem 265(8): 4290-5.<br />

Nakagawa, M., T. Miyamoto, et al. (2001). "O-Glycosylation <strong>of</strong> G-protein-coupled receptor, octopus rhodopsin. Direct<br />

analysis by FAB mass spectrometry." FEBS Lett 496(1): 19-24.<br />

Neddermann, P., C. Gargioli, et al. (2003). "A novel, inducible, eukaryotic gene expression system based on <strong>the</strong><br />

quorum-sensing transcription factor TraR." EMBO Rep 4(2): 159-65.<br />

Nekrasova, E., A. Sosinskaya, et al. (1996). "Overexpression, solubilization and purification <strong>of</strong> rat and human olfactory<br />

receptors." Eur J Biochem 238(1): 28-37.<br />

Nicholls, D. J., Tomkinsson, N.P., Wiley, K.E., Brammal, A., Bowers, L., Grahames, C., Gaw, A., Meghani, P., shelton, P.,<br />

Wright, T.J., Mallinder, P.R. (2008). "Identification <strong>of</strong> a putative intracellular allosteric antagonist binding site<br />

in <strong>the</strong> CXC chemokine receptors 1 and 2." Mol Pharmacol 74: 1193-1202.<br />

Niebauer, R. T. and A. S. Robinson (2006). "Exceptional total and functional yields <strong>of</strong> <strong>the</strong> human adenosine (A2a)<br />

receptor expressed in <strong>the</strong> yeast Saccharomyces cerevisiae." Protein Expr Purif 46(2): 204-11.<br />

Nomura, E., K. Katsuta, et al. (2004). "Acid-labile surfactant improves in-sodium dodecyl sulfate polyacrylamide gel<br />

protein digestion for matrix-assisted laser desorption/ionization mass spectrometric peptide mapping." J<br />

Mass Spectrom 39(2): 202-7.<br />

Nonaka, H., S. Otaki, et al. (1998). "Unique binding pocket for KW-4679 in <strong>the</strong> histamine H1 receptor." Eur J Pharmacol<br />

345(1): 111-7.<br />

Nygaard, R., T. M. Frimurer, et al. (2009). "Ligand binding and micro-switches in 7TM receptor structures." Trends<br />

Pharmacol Sci 30(5): 249-59.<br />

O'Malley, M. A., T. Lazarova, et al. (2007). "High-level expression in Saccharomyces cerevisiae enables isolation and<br />

spectroscopic characterization <strong>of</strong> functional human adenosine A2a receptor." J Struct Biol 159(2): 166-78.<br />

Obosi, L. A., D. G. Schuette, et al. (1996). "Functional characterisation <strong>of</strong> <strong>the</strong> Drosophila 5-HTdro1 and 5-HTdro2B<br />

serotonin receptors in insect cells: activation <strong>of</strong> a G(alpha) s-like protein by 5-HTdro1 but lack <strong>of</strong> coupling to<br />

inhibitory G-proteins by 5-HTdro2B." FEBS Lett 381(3): 233-6.<br />

Offermanns, S. (2003). "G-proteins as transducers in transmembrane signalling." Prog Biophys Mol Biol 83(2): 101-<br />

30.<br />

Ogonah, O. W., R. B. Freedman, et al. (1995). "Analysis <strong>of</strong> human interferon-gamma glyc<strong>of</strong>orms produced in<br />

baculovirus infected insect cells by matrix assisted laser desorption spectrometry." Biochem Soc Trans<br />

23(1): 100S.<br />

Ohta, K., H. Hayashi, et al. (1994). "Site-directed mutagenesis <strong>of</strong> <strong>the</strong> histamine H1 receptor: roles <strong>of</strong> aspartic acid107,<br />

asparagine198 and threonine194." Biochem Biophys Res Commun 203(2): 1096-101.<br />

Ohtaki, T., K. Ogi, et al. (1998). "Expression, purification, and reconstitution <strong>of</strong> receptor for pituitary adenylate cyclaseactivating<br />

polypeptide. large-scale purification <strong>of</strong> a functionally active G protein-coupled receptor produced<br />

in Sf9 insect cells." J Biol Chem 273(25): 15464-73.<br />

Ollivon, M., O. Eidelman, et al. (1988). "Micelle-vesicle transition <strong>of</strong> egg phosphatidylcholine and octyl glucoside."<br />

Biochemistry 27(5): 1695-703.<br />

Opekarova, M. and W. Tanner (2003). "Specific lipid requirements <strong>of</strong> membrane proteins--a putative bottleneck in<br />

heterologous expression." Biochim Biophys Acta 1610(1): 11-22.<br />

Osband, M. and R. McCaffrey (1979). "Solubilization, separation, and partial characterization <strong>of</strong> histamine H1 and H2<br />

receptors from calf thymocyte membranes." J Biol Chem 254(20): 9970-2.<br />

Palczewski, K., T. Kumasaka, et al. (2000). "Crystal structure <strong>of</strong> rhodopsin: A G protein-coupled receptor." Science<br />

289(5480): 739-45.<br />

141 | P a g e


References<br />

Palva, I., P. Lehtovaara, et al. (1983). "Secretion <strong>of</strong> interferon by Bacillus subtilis." Gene 22(2-3): 229-35.<br />

Panneels, V., C. Eroglu, et al. (2003). "Pharmacological characterization and immunoaffinity purification <strong>of</strong><br />

metabotropic glutamate receptor from Drosophila overexpressed in Sf9 cells." Protein Expr Purif 30(2):<br />

275-82.<br />

Pardo, L., X. Deupi, et al. (2007). "The Role <strong>of</strong> Internal Water Molecules in <strong>the</strong> Structure and Function <strong>of</strong> <strong>the</strong> Rhodopsin<br />

Family <strong>of</strong> G Protein-Coupled Receptors." Chembiochem 8(1): 19-24.<br />

Park, J. H., P. Scheerer, et al. (2008). "Crystal structure <strong>of</strong> <strong>the</strong> ligand-free G-protein-coupled receptor opsin." Nature<br />

454(7201): 183-7.<br />

Park, P. S. and J. W. Wells (2003). "Monomers and oligomers <strong>of</strong> <strong>the</strong> M2 muscarinic cholinergic receptor purified from<br />

Sf9 cells." Biochemistry 42(44): 12960-71.<br />

Parker, E. M., K. Kameyama, et al. (1991). "Reconstitutively active G protein-coupled receptors purified from<br />

baculovirus-infected insect cells." J Biol Chem 266(1): 519-27.<br />

Patel, A. B., E. Crocker, et al. (2005). "Changes in interhelical hydrogen bonding upon rhodopsin activation." J. Mol. Biol.<br />

347(4): 803-12.<br />

Paulsen, H., B. Finkenzeller, et al. (1993). "Pigments induce folding <strong>of</strong> light-harvesting chlorophyll a/b-binding<br />

protein." Eur J Biochem 215(3): 809-16.<br />

Pellissier, L., J. Sallander, et al. (2009). "Conformational toggle switches implicated in basal constitutive and agonistinduced<br />

activated states <strong>of</strong> 5-HT4 receptors." Mol. Pharmacol. 75: 982-990.<br />

Pierce, K. L., R. T. Premont, et al. (2002). "Seven-transmembrane receptors." Nat Rev Mol Cell Biol 3(9): 639-50.<br />

Pornillos, O., Y. J. Chen, et al. (2005). "X-ray structure <strong>of</strong> <strong>the</strong> EmrE multidrug transporter in complex with a substrate."<br />

Science 310(5756): 1950-3.<br />

Rands, E., M. R. Candelore, et al. (1990). "Mutational analysis <strong>of</strong> beta-adrenergic receptor glycosylation." J Biol Chem<br />

265(18): 10759-64.<br />

Rappen-Cremer, E., Borchard, U., Hafner, D., Berger, F. (1989). "H1 receptor reserves in guinea pig left atria, trachea<br />

and pig coronary artery as identified by phenoxybenzamine." Agents Actions 28: 218-223.<br />

Rasmussen, S. G., H. J. Choi, et al. (2007). "Crystal structure <strong>of</strong> <strong>the</strong> human beta2 adrenergic G-protein-coupled<br />

receptor." Nature 450(7168): 383-7.<br />

Ratnala, V. R., S. R. Kiihne, et al. (2007). "Solid-state NMR evidence for a protonation switch in <strong>the</strong> binding pocket <strong>of</strong> <strong>the</strong><br />

H1 receptor upon binding <strong>of</strong> <strong>the</strong> agonist histamine." J Am Chem Soc 129(4): 867-72.<br />

Ratnala, V. R., H. G. Swarts, et al. (2004). "Large-scale overproduction, functional purification and ligand affinities <strong>of</strong> <strong>the</strong><br />

His-tagged human histamine H1 receptor." Eur J Biochem 271(13): 2636-46.<br />

Reeves, P. J., N. Callewaert, et al. (2002). "Structure and function in rhodopsin: high-level expression <strong>of</strong> rhodopsin with<br />

restricted and homogeneous N-glycosylation by a tetracycline-inducible N-acetylglucosaminyltransferase Inegative<br />

HEK293S stable mammalian cell line." Proc Natl Acad Sci U S A 99(21): 13419-24.<br />

Reeves, P. J., J. M. Kim, et al. (2002). "Structure and function in rhodopsin: a tetracycline-inducible system in stable<br />

mammalian cell lines for high-level expression <strong>of</strong> opsin mutants." Proc Natl Acad Sci U S A 99(21): 13413-8.<br />

Reilander, H., F. Boege, et al. (1991). "Purification and functional characterization <strong>of</strong> <strong>the</strong> human beta 2-adrenergic<br />

receptor produced in baculovirus-infected insect cells." FEBS Lett 282(2): 441-4.<br />

Reiter, E. and R. J. Lefkowitz (2006). "GRKs and beta-arrestins: roles in receptor silencing, trafficking and signaling."<br />

Trends Endocrinol Metab 17(4): 159-65.<br />

Remigy, H. W., D. Caujolle-Bert, et al. (2003). "Membrane protein reconstitution and crystallization by controlled<br />

dilution." FEBS Lett 555(1): 160-9.<br />

Rens-Domiano, S. and T. Reisine (1991). "Structural analysis and functional role <strong>of</strong> <strong>the</strong> carbohydrate component <strong>of</strong><br />

somatostatin receptors." J Biol Chem 266(30): 20094-102.<br />

Rhee, S. G. (2001). "Regulation <strong>of</strong> phosphoinositide-specific phospholipase C." Annu Rev Biochem 70: 281-312.<br />

Riek, R. P., A. A. Finch, et al. (2008). "Wide turn diversity in protein transmembrane helices implications for G-proteincoupled<br />

receptor and o<strong>the</strong>r polytopic membrane protein structure and function." Mol. Pharmacol. 73(4):<br />

1092-104.<br />

Rigaud, J. L., G. Mosser, et al. (1997). "Bio-Beads: an efficient strategy for two-dimensional crystallization <strong>of</strong> membrane<br />

proteins." J Struct Biol 118(3): 226-35.<br />

Ringkananont, U., J. Van Durme, et al. (2006). "Repulsive separation <strong>of</strong> <strong>the</strong> cytoplasmic ends <strong>of</strong> transmembrane helices<br />

3 and 6 is linked to receptor activation in a novel thyrotropin receptor mutant (M626I)." Mol Endocrinol<br />

20(4): 893-903.<br />

Roosild, T. P., J. Greenwald, et al. (2005). "NMR structure <strong>of</strong> Mistic, a membrane-integrating protein for membrane<br />

protein expression." Science 307(5713): 1317-21.<br />

142 | P a g e


References<br />

Rosenbaum, D. M., S. G. Rasmussen, et al. (2009). "The structure and function <strong>of</strong> G-protein-coupled receptors." Nature<br />

459(7245): 356-63.<br />

Rovati, G. E., V. Capra, et al. (2007). "The highly conserved DRY motif <strong>of</strong> class A G protein-coupled receptors: beyond<br />

<strong>the</strong> ground state." Mol Pharmacol 71(4): 959-64.<br />

Roy, A., A. K. Shukla, et al. (2008). "Employing Rhodobacter sphaeroides to functionally express and purify human G<br />

protein-coupled receptors." Biol Chem 389(1): 69-78.<br />

Ruan, K. H., V. Cervantes, et al. (2008). "A simple, quick, and high-yield preparation <strong>of</strong> <strong>the</strong> human thromboxane A2<br />

receptor in full size for structural studies." Biochemistry 47(26): 6819-26.<br />

Ruprecht, J. J., T. Mielke, et al. (2004). "Electron crystallography reveals <strong>the</strong> structure <strong>of</strong> metarhodopsin I." Embo J<br />

23(18): 3609-20.<br />

Russo, D., G. D. Chazenbalk, et al. (1991). "Site-directed mutagenesis <strong>of</strong> <strong>the</strong> human thyrotropin receptor: role <strong>of</strong><br />

asparagine-linked oligosaccharides in <strong>the</strong> expression <strong>of</strong> a functional receptor." Mol Endocrinol 5(1): 29-33.<br />

Sah, V. P., T. M. Seasholtz, et al. (2000). "The role <strong>of</strong> Rho in G protein-coupled receptor signal transduction." Annu Rev<br />

Pharmacol Toxicol 40: 459-89.<br />

Salom, D., D. T. Lodowski, et al. (2006). "Crystal structure <strong>of</strong> a photoactivated deprotonated intermediate <strong>of</strong><br />

rhodopsin." Proc Natl Acad Sci U S A 103(44): 16123-8.<br />

Salom, D., N. Wu, et al. (2008). "Heterologous expression and purification <strong>of</strong> <strong>the</strong> serotonin type 4 receptor from<br />

transgenic mouse retina." Biochemistry 47(50): 13296-307.<br />

Sander, P., S. Grunewald, et al. (1994). "Constitutive expression <strong>of</strong> <strong>the</strong> human D2S-dopamine receptor in <strong>the</strong><br />

unicellular yeast Saccharomyces cerevisiae." Biochim Biophys Acta 1193(2): 255-62.<br />

Sansuk, K., C. I. Balog, et al. (2008). "GPCR Proteomics: Mass Spectrometric and Functional Analysis <strong>of</strong> Histamine H1<br />

Receptor after Baculovirus-Driven and in Vitro Cell Free Expression." J Proteome Res.<br />

Sarramegn, V., I. Muller, et al. (2006). "Recombinant G protein-coupled receptors from expression to renaturation: a<br />

challenge towards structure." Cell Mol Life Sci 63(10): 1149-64.<br />

Sarramegna, V., I. Muller, et al. (2005). "Solubilization, purification, and mass spectrometry analysis <strong>of</strong> <strong>the</strong> human muopioid<br />

receptor expressed in Pichia pastoris." Protein Expr Purif 43(2): 85-93.<br />

Sarramegna, V., F. Talmont, et al. (2003). "Heterologous expression <strong>of</strong> G-protein-coupled receptors: comparison <strong>of</strong><br />

expression systems fron <strong>the</strong> standpoint <strong>of</strong> large-scale production and purification." Cell Mol Life Sci 60(8):<br />

1529-46.<br />

Satoh, H. and J. Inui (1984). "Endo<strong>the</strong>lial cell-dependent relaxation and contraction induced by histamine in <strong>the</strong><br />

isolated guinea-pig pulmonary artery." Eur J Pharmacol 97(3-4): 321-4.<br />

Sawasaki, T., T. Ogasawara, et al. (2002). "A cell-free protein syn<strong>the</strong>sis system for high-throughput proteomics." Proc<br />

Natl Acad Sci U S A 99(23): 14652-7.<br />

Scheer, A., F. Fanelli, et al. (1996). "Constitutively active mutants <strong>of</strong> <strong>the</strong> alpha 1B-adrenergic receptor: role <strong>of</strong> highly<br />

conserved polar amino acids in receptor activation." EMBO J. 15(14): 3566-78.<br />

Scheerer, P., J. H. Park, et al. (2008). "Crystal structure <strong>of</strong> opsin in its G-protein-interacting conformation." Nature<br />

455(7212): 497-502.<br />

Schiller, H., E. Molsberger, et al. (2001). "Solubilization and purification <strong>of</strong> <strong>the</strong> human ETB endo<strong>the</strong>lin receptor<br />

produced by high-level fermentation in Pichia pastoris." Receptors Channels 7(6): 453-69.<br />

Schipani, E., K. Kruse, et al. (1995). "A constitutively active mutant PTH-PTHrP receptor in Jansen-type metaphyseal<br />

chondrodysplasia." Science 268(5207): 98-100.<br />

Schlyer, S. and R. Horuk (2006). "I want a new drug: G-protein-coupled receptors in drug development." Drug Discov<br />

Today 11(11-12): 481-93.<br />

Scott, M. J., S. S. Modha, et al. (2007). "Efficient expression <strong>of</strong> secreted proteases via recombinant BacMam virus."<br />

Protein Expr Purif 52(1): 104-16.<br />

Seddon, A. M., P. Curnow, et al. (2004). "Membrane proteins, lipids and detergents: not just a soap opera." Biochim<br />

Biophys Acta 1666(1-2): 105-17.<br />

Seifert, R., L. Grunbaum, et al. (1994). "Histamine H1-receptors in HL-60 monocytes are coupled to Gi-proteins and<br />

pertussis toxin-insensitive G-proteins and mediate activation <strong>of</strong> Ca2+ influx without concomitant Ca2+<br />

mobilization from intracellular stores." Naunyn Schmiedebergs Arch Pharmacol 349(4): 355-61.<br />

Sen, S., V. P. Jaakola, et al. (2003). "Functional expression and direct visualization <strong>of</strong> <strong>the</strong> human alpha 2B -adrenergic<br />

receptor and alpha 2B -AR-green fluorescent fusion protein in mammalian cell using Semliki Forest virus<br />

vectors." Protein Expr Purif 32(2): 265-75.<br />

Serrano-Vega, M. J., F. Magnani, et al. (2008). "Conformational <strong>the</strong>rmostabilization <strong>of</strong> <strong>the</strong> beta1-adrenergic receptor in<br />

a detergent-resistant form." Proc Natl Acad Sci U S A 105(3): 877-82.<br />

143 | P a g e


References<br />

Shapiro, D. A., K. Kristiansen, et al. (2002). "Evidence for a model <strong>of</strong> agonist-induced activation <strong>of</strong> 5hydroxytryptamine<br />

2A serotonin receptors that involves <strong>the</strong> disruption <strong>of</strong> a strong ionic interaction between<br />

helices 3 and 6." J Biol Chem 277(13): 11441-9.<br />

Shenker, A., L. Laue, et al. (1993). "A constitutively activating mutation <strong>of</strong> <strong>the</strong> luteinizing hormone receptor in familial<br />

male precocious puberty." Nature 365(6447): 652-4.<br />

Shi, L., G. Liapakis, et al. (2002). "Beta2 adrenergic receptor activation. Modulation <strong>of</strong> <strong>the</strong> proline kink in<br />

transmembrane 6 by a rotamer toggle switch." J Biol Chem 277(43): 40989-96.<br />

Shimamura, T., K. Hiraki, et al. (2008). "Crystal structure <strong>of</strong> squid rhodopsin with intracellularly extended cytoplasmic<br />

region." J Biol Chem 283(26): 17753-6.<br />

Shoji, I., H. Aizaki, et al. (1997). "Efficient gene transfer into various mammalian cells, including non-hepatic cells, by<br />

baculovirus vectors." J Gen Virol 78 ( Pt 10): 2657-64.<br />

Shulman-Roskes, E. M., D. A. Noe, et al. (1998). "The partitioning <strong>of</strong> phosphoramide mustard and its aziridinium ions<br />

among alkylation and P-N bond hydrolysis reactions." J Med Chem 41(4): 515-29.<br />

Siehler, S. (2008). "Cell-based assays in GPCR drug discovery." Biotechnol J 3(4): 471-83.<br />

Sinz, A. (2007). "Investigation <strong>of</strong> protein-ligand interactions by mass spectrometry." ChemMedChem 2(4): 425-31.<br />

Smit, M. J., H. F. Vischer, et al. (2007). "Pharmacogenomic and structural analysis <strong>of</strong> constitutive g protein-coupled<br />

receptor activity." Annu Rev Pharmacol Toxicol 47: 53-87.<br />

Smith, G. E., M. D. Summers, et al. (1983). "Production <strong>of</strong> human beta interferon in insect cells infected with a<br />

baculovirus expression vector." Mol Cell Biol 3(12): 2156-65.<br />

Son, C. D., H. Sargsyan, et al. (2005). "Analysis <strong>of</strong> ligand-receptor cross-linked fragments by mass spectrometry." J Pept<br />

Res 65(3): 418-26.<br />

Spalding, T. A., E. S. Burstein, et al. (1998). "Identification <strong>of</strong> a ligand-dependent switch within a muscarinic receptor." J<br />

Biol Chem 273(34): 21563-8.<br />

Spiegel, A. M. (1998). G Proteins, Receptors and Diseases. A. M. Spiegel. Totawa, Humana Press: 1-22.<br />

Spirin, A. S. (2004). "High-throughput cell-free systems for syn<strong>the</strong>sis <strong>of</strong> functionally active proteins." Trends<br />

Biotechnol 22(10): 538-45.<br />

Spirin, A. S., V. I. Baranov, et al. (1988). "A continuous cell-free translation system capable <strong>of</strong> producing polypeptides in<br />

high yield." Science 242(4882): 1162-4.<br />

Springael, J. Y., C. de Poorter, et al. (2007). "The activation mechanism <strong>of</strong> chemokine receptor CCR5 involves common<br />

structural changes but a different network <strong>of</strong> interhelical interactions relative to rhodopsin." Cell. Signal.<br />

19(7): 1446-1456.<br />

Staedtler, P., S. Hoenig, et al. (1995). "Identification <strong>of</strong> <strong>the</strong> active-site nucleophile in 6-phospho-beta-galactosidase<br />

from Staphylococcus aureus by labelling with syn<strong>the</strong>tic inhibitors." Eur J Biochem 232(2): 658-63.<br />

Stanasila, L., F. Pattus, et al. (1998). "Heterologous expression <strong>of</strong> G-protein-coupled receptors: human opioid receptors<br />

under scrutiny." Biochimie 80(5-6): 563-71.<br />

Standfuss, J., G. Xie, et al. (2007). "Crystal structure <strong>of</strong> a <strong>the</strong>rmally stable rhodopsin mutant." J Mol Biol 372(5): 1179-<br />

88.<br />

Stenkamp, R. E., S. Filipek, et al. (2002). "Crystal structure <strong>of</strong> rhodopsin: a template for cone visual pigments and o<strong>the</strong>r<br />

G protein-coupled receptors." Biochim Biophys Acta 1565(2): 168-82.<br />

Strange, P. G. (2002). "Mechanisms <strong>of</strong> inverse agonism at G-protein-coupled receptors." Trends Pharmacol Sci 23(2):<br />

89-95.<br />

Stru<strong>the</strong>rs, M., H. Yu, et al. (2000). "G protein-coupled receptor activation: analysis <strong>of</strong> a highly constrained,<br />

"straitjacketed" rhodopsin." Biochemistry 39(27): 7938-42.<br />

Takagi, K. and M. Uchida (1970). "Purification <strong>of</strong> histamine receptor in cat small intestinal muscle. I. Specificity <strong>of</strong><br />

labeling histamine receptor with radioactive dibenamine." Jpn J Pharmacol 20(2): 272-86.<br />

Takayanagi, I., Satoh, M., Shinkai, M., Okayasu, A., hosoki, R. (1996). "Phenoxybenzamine discriminates between two<br />

distinct populations <strong>of</strong> histamine H1 receptors in longitudinal muscle <strong>of</strong> guinea pig ileum." J Smooth Muscle<br />

Res 32: 249-254.<br />

Talmont, F., S. Sidobre, et al. (1996). "Expression and pharmacological characterization <strong>of</strong> <strong>the</strong> human mu-opioid<br />

receptor in <strong>the</strong> methylotrophic yeast Pichia pastoris." FEBS Lett 394(3): 268-72.<br />

Tate, C. G. (2001). "Overexpression <strong>of</strong> mammalian integral membrane proteins for structural studies." FEBS Lett<br />

504(3): 94-8.<br />

Teller, D. C., T. Okada, et al. (2001). "Advances in determination <strong>of</strong> a high-resolution three-dimensional structure <strong>of</strong><br />

rhodopsin, a model <strong>of</strong> G-protein-coupled receptors (GPCRs)." Biochemistry 40(26): 7761-72.<br />

144 | P a g e


References<br />

ter Laak, A. M., H. Timmerman, et al. (1995). "Modelling and mutation studies on <strong>the</strong> histamine H1-receptor agonist<br />

binding site reveal different binding modes for H1-agonists: Asp116 (TM3) has a constitutive role in receptor<br />

stimulation." J Comput Aided Mol Des 9(4): 319-30.<br />

Terner, U. K., Cook, D.A., Mark, G.S. (1971). "Studies <strong>of</strong> <strong>the</strong> chemical nature <strong>of</strong> <strong>the</strong> alpha adrenergic receptor-V. Validity<br />

<strong>of</strong> <strong>the</strong> receptor protection approach." Biochem Pharmacol 20: 597-603.<br />

Thurmond, R. L., E. W. Gelfand, et al. (2008). "The role <strong>of</strong> histamine H1 and H4 receptors in allergic inflammation: <strong>the</strong><br />

search for new antihistamines." Nat Rev Drug Discov 7(1): 41-53.<br />

Toll, L. and S. H. Snyder (1982). "Solubilization and Characterization <strong>of</strong> Histamine H1 receptors in brain." J Biol Chem<br />

257(22): 13593-601.<br />

Topiol, S. and M. Sabio (2008). "Use <strong>of</strong> <strong>the</strong> X-ray structure <strong>of</strong> <strong>the</strong> Beta2-adrenergic receptor for drug discovery." Bioorg<br />

Med Chem Lett 18(5): 1598-602.<br />

Topiol, S. and M. Sabio (2009). "X-ray structure breakthroughs in <strong>the</strong> GPCR transmembrane region." Biochem<br />

Pharmacol 78(1): 11-20.<br />

Tucker, J. and R. Grisshammer (1996). "Purification <strong>of</strong> a rat neurotensin receptor expressed in Escherichia coli."<br />

Biochem J 317 ( Pt 3): 891-9.<br />

Turner, G. J., R. Reusch, et al. (1999). "Heterologous gene expression in a membrane-protein-specific system." Protein<br />

Expr Purif 17(2): 312-23.<br />

Urizar, E., S. Claeysen, et al. (2005). "An activation switch in <strong>the</strong> rhodopsin family <strong>of</strong> G protein-coupled receptors: <strong>the</strong><br />

thyrotropin receptor." J Biol Chem 280(17): 17135-41.<br />

van Montfort, B. A., B. Canas, et al. (2002). "Improved in-gel approaches to generate peptide maps <strong>of</strong> integral<br />

membrane proteins with matrix-assisted laser desorption/ionization time-<strong>of</strong>-flight mass spectrometry." J<br />

Mass Spectrom 37(3): 322-30.<br />

van Montfort, B. A., M. K. Doeven, et al. (2002). "Combined in-gel tryptic digestion and CNBr cleavage for <strong>the</strong><br />

generation <strong>of</strong> peptide maps <strong>of</strong> an integral membrane protein with MALDI-TOF mass spectrometry." Biochim<br />

Biophys Acta 1555(1-3): 111-5.<br />

van Rijn, R. M. (2007). Molecular pharmacology <strong>of</strong> <strong>the</strong> human histamine H4 receptor. Medicinal chemistry.<br />

<strong>Amsterdam</strong>, <strong>Vrije</strong> <strong>Universiteit</strong> <strong>Amsterdam</strong>: 159-176.<br />

Vanhauwe, J. F., T. O. Thomas, et al. (2002). "Thrombin receptors activate G(o) proteins in endo<strong>the</strong>lial cells to regulate<br />

intracellular calcium and cell shape changes." J Biol Chem 277(37): 34143-9.<br />

Varga, C., K. Horvath, et al. (2005). "Inhibitory effects <strong>of</strong> histamine H4 receptor antagonists on experimental colitis in<br />

<strong>the</strong> rat." Eur J Pharmacol.<br />

Vicentic, A., A. Robeva, et al. (2002). "Biochemistry and pharmacology <strong>of</strong> epitope-tagged alpha(1)-adrenergic receptor<br />

subtypes." J Pharmacol Exp Ther 302(1): 58-65.<br />

Vilardaga, J. P. (2006). "Switching modes for G protein-coupled receptor activation." Nat Chem Biol 2(8): 395-6.<br />

Wagner, S., M. L. Bader, et al. (2006). "Rationalizing membrane protein overexpression." Trends Biotechnol 24(8):<br />

364-71.<br />

Walsh, M. T., J. F. Foley, et al. (1998). "Characterization <strong>of</strong> <strong>the</strong> role <strong>of</strong> N-linked glycosylation on <strong>the</strong> cell signaling and<br />

expression <strong>of</strong> <strong>the</strong> human thromboxane A2 receptor alpha and beta is<strong>of</strong>orms." J Pharmacol Exp Ther 286(2):<br />

1026-36.<br />

Warne, T., J. Chirnside, et al. (2003). Expression and purification <strong>of</strong> truncated, non-glycosylated turkey betaadrenergic<br />

receptors for crystallization. Biochim Biophys Acta. 1610: 133-40.<br />

Warne, T., M. J. Serrano-Vega, et al. (2008). "Structure <strong>of</strong> a beta1-adrenergic G-protein-coupled receptor." Nature<br />

454(7203): 486-91.<br />

Weber, W., M. Rimann, et al. (2004). "Gas-inducible transgene expression in mammalian cells and mice." Nat<br />

Biotechnol 22(11): 1440-4.<br />

Weis, W. I. and B. K. Kobilka (2008). "Structural insights into G-protein-coupled receptor activation." Curr. Opin. Struct.<br />

Biol. 18(6): 734-40.<br />

Weiss, H. M. and R. Grisshammer (2002). "Purification and characterization <strong>of</strong> <strong>the</strong> human adenosine A(2a) receptor<br />

functionally expressed in Escherichia coli." Eur J Biochem 269(1): 82-92.<br />

Wheatley, M. and S. R. Hawtin (1999). "Glycosylation <strong>of</strong> G-protein-coupled receptors for hormones central to normal<br />

reproductive functioning: its occurrence and role." Hum Reprod Update 5(4): 356-64.<br />

White, J. F., L. B. Trinh, et al. (2004). "Automated large-scale purification <strong>of</strong> a G protein-coupled receptor for<br />

neurotensin." FEBS Lett 564(3): 289-93.<br />

Wieland, K., A. M. Laak, et al. (1999). "Mutational analysis <strong>of</strong> <strong>the</strong> antagonist-binding site <strong>of</strong> <strong>the</strong> histamine H(1)<br />

receptor." J Biol Chem 274(42): 29994-30000.<br />

145 | P a g e


References<br />

Wuhrer, M., M. I. Catalina, et al. (2007). "Glycoproteomics based on tandem mass spectrometry <strong>of</strong> glycopeptides." J<br />

Chromatogr B Analyt Technol Biomed Life Sci 849(1-2): 115-28.<br />

Xu, W., S. A. Filppula, et al. (2005). "Purification and mass spectroscopic analysis <strong>of</strong> human CB1 cannabinoid receptor<br />

functionally expressed using <strong>the</strong> baculovirus system." J Pept Res 66(3): 138-50.<br />

Yabuki, T., T. Kigawa, et al. (1998). "Dual amino acid-selective and site-directed stable-isotope labeling <strong>of</strong> <strong>the</strong> human c-<br />

Ha-Ras protein by cell-free syn<strong>the</strong>sis." J Biomol NMR 11(3): 295-306.<br />

Yao, X., C. Parnot, et al. (2006). "Coupling ligand structure to specific conformational switches in <strong>the</strong> beta2adrenoceptor."<br />

Nat Chem Biol 2(8): 417-22.<br />

Yeliseev, A. A., K. K. Wong, et al. (2005). "Expression <strong>of</strong> human peripheral cannabinoid receptor for structural studies."<br />

Protein Sci 14(10): 2638-53.<br />

Yu, Y. Q., M. Gilar, et al. (2004). "A complete peptide mapping <strong>of</strong> membrane proteins: a novel surfactant aiding <strong>the</strong><br />

enzymatic digestion <strong>of</strong> bacteriorhodopsin." Rapid Commun Mass Spectrom 18(6): 711-5.<br />

Yu, Y. Q., M. Gilar, et al. (2003). "Enzyme-friendly, mass spectrometry-compatible surfactant for in-solution enzymatic<br />

digestion <strong>of</strong> proteins." Anal Chem 75(21): 6023-8.<br />

Zhang, L., D. Salom, et al. (2005). "Expression <strong>of</strong> functional G protein-coupled receptors in photoreceptors <strong>of</strong><br />

transgenic Xenopus laevis." Biochemistry 44(44): 14509-18.<br />

Zhang, R., T. K. Kim, et al. (2007). "Biochemical and mass spectrometric characterization <strong>of</strong> <strong>the</strong> human CB2<br />

cannabinoid receptor expressed in Pichia pastoris-Importance <strong>of</strong> correct processing <strong>of</strong> <strong>the</strong> N-terminus."<br />

Protein Expr Purif.<br />

Zhang, Z., S. C. Austin, et al. (2001). "Glycosylation <strong>of</strong> <strong>the</strong> human prostacyclin receptor: role in ligand binding and signal<br />

transduction." Mol Pharmacol 60(3): 480-7.<br />

Zingel, V., C. Leschke, et al. (1995). "Developments in histamine H1-receptor agonists." Prog Drug Res 44: 49-85.<br />

Zvonok, N., S. Yaddanapudi, et al. (2007). "Comprehensive proteomic mass spectrometric characterization <strong>of</strong> human<br />

cannabinoid CB2 receptor." J Proteome Res 6(6): 2068-79.<br />

146 | P a g e


Acknowledgements<br />

Jib<br />

[Type <strong>the</strong> company name]<br />

[Pick <strong>the</strong> date]


Because <strong>the</strong>re is no “I” in “TEAM”…<br />

My Ph.D. years at <strong>the</strong> <strong>Vrije</strong> <strong>Universiteit</strong> and this<br />

<strong>the</strong>sis have had mentorship from numerous<br />

individuals both from within <strong>the</strong> university and<br />

outside <strong>of</strong> it. It is to <strong>the</strong>se individuals that my<br />

heartfelt gratitude and thanks go out to, for<br />

without <strong>the</strong>ir help, this <strong>the</strong>sis would not have<br />

seen its ISBN number.<br />

The supervising team<br />

My promoter, Pr<strong>of</strong>. Dr. Rob Leurs, thank you for<br />

giving me a chance to do my Ph.D. study in your group.<br />

Your knowledge and logical way <strong>of</strong> thinking have been<br />

<strong>of</strong> great value for me and have provided a good<br />

structure for this <strong>the</strong>sis. I also would like to express<br />

my gratitude to my co-promoter, Dr. Remko Bakker,<br />

thank you for your guidance and your promptly<br />

responses to my questions, drafts and email through<br />

<strong>the</strong>se years. You are <strong>the</strong> one who put <strong>the</strong> paper from<br />

Ishihara (<strong>the</strong> first paper showing that GPCRs can be<br />

expressed by cell free expression system) in my<br />

mailbox and 3 years later, we finally made it happen<br />

with our H1R protein. I am still keeping that bottle <strong>of</strong><br />

champagne you gave me as <strong>the</strong> trophy <strong>of</strong> our victory.<br />

This work would not have been possible without Pr<strong>of</strong>.<br />

Dr. Wim de Grip, Dear Wim, your expertise and<br />

connections in membrane protein biochemistry and<br />

biophysics are very valuable and I am so fortunate to<br />

be supervised by you. My thanks also go to <strong>the</strong><br />

members at <strong>the</strong> NCMLS, Nijmegen. Arthur, thank you<br />

for helping me with FT-IR. Jenny, thank you for<br />

teaching me all <strong>the</strong> techniques related to BVES. And<br />

very special thanks to Petra, Dear Petra; besides<br />

showing me how to purify membrane proteins, you<br />

are also my guardian angel. Thank you for making me<br />

feels like home in Nijmegen. I deeply enjoyed your<br />

company and a free-pass to <strong>the</strong> pit in Assen.<br />

I would like to thank Dr. Paul Hensbergen and his<br />

team Crina and Anne for <strong>the</strong>ir contribution in chapter<br />

3 and 4 in this <strong>the</strong>sis. Dear Paul, you are my MS guru. I<br />

really enjoyed my visits to Leiden and our telephone<br />

conversations. As much as you can be very serious,<br />

148 | P a g e<br />

you can also laugh very easily. It really helps making<br />

such a good vibe in our collaboration.<br />

I am also indebted to Pr<strong>of</strong>. Dr. Leonardo Pardo and<br />

his team Dr. Xavier Deupi and Ivan Torrecillas for<br />

<strong>the</strong>ir contribution in chapter 5 and 6 in this <strong>the</strong>sis.<br />

Without <strong>the</strong>ir help in structure modeling, I probably<br />

would not be able to solve <strong>the</strong> puzzles regarding our<br />

H1R mutants.<br />

I wish to express my thank to Pr<strong>of</strong>. Dr. Raymond<br />

Booth, whose valuable comments and funding has<br />

been <strong>of</strong> great support to our work. My thanks also go<br />

to <strong>the</strong> reading committee whose name has not been<br />

mentioned yet. Thank you Pr<strong>of</strong>. Dr. Martine Smit, Dr.<br />

Astrid Alewijnse and Dr. Marco Siderius for<br />

reviewing my <strong>the</strong>sis. Also many thanks to my former<br />

colleagues, Dr. Martijn Bruyster and Dr. Aldo<br />

Jongejan for <strong>the</strong>ir valuable comments and advice<br />

regarding my work with <strong>the</strong> H1R.<br />

The scientific supporters<br />

Richard van Rijn, my dear friend, your multi-tasking<br />

and self-discipline has never failed to impress and<br />

inspire me. Thank you for having faith in me, for<br />

always encouraging and keeping me focused, for<br />

spending hours cheering me up through <strong>the</strong> phone<br />

lines, for being <strong>the</strong>re through my good and bad times,<br />

for letting me be your friend, for staying friend with<br />

me and for being <strong>the</strong> best, best friend in <strong>the</strong> world.<br />

Gerold Bongers, thank you for always taking me<br />

along on your motorbike. For you it is adrenaline<br />

kicks, for me it is meditations (since I manage to fall<br />

asleep anyway). I also appreciate that you show up at<br />

every running races. I probable never said it before<br />

but you are <strong>the</strong> best supporters and definitely<br />

deserved some credits for those medals.<br />

Laura de Wit, You always greet me with a warm smile<br />

and open arms every time I dropped by your <strong>of</strong>fice.<br />

Thank you for your friendship and kind helps in many<br />

different ways.<br />

Andre van Marle and Tonio (JOSÉ ANTONIO ARIAS<br />

MONTAÑO), thank you for all your helps and


comments, especially in my first year. Dear Andre, <strong>the</strong><br />

book “Cell Free expression system” that you gave me<br />

will always be my holy bible. Dear Tonio, what you<br />

taught me on my first week is a good basis <strong>of</strong> my<br />

purification work up till now. Also, thank you for <strong>the</strong><br />

lovely present from Mexico. (Although, it is funny to<br />

see that o<strong>the</strong>r people in <strong>the</strong> lab got bottles <strong>of</strong> tequila,<br />

but I got a Barbie doll.)<br />

Herman Lim and Atilla Akdemir, thank you for your<br />

friendship and patience when listening to all my<br />

miserable stories. Dear Herman, I enjoyed <strong>the</strong> time we<br />

shared <strong>the</strong> <strong>of</strong>fice toge<strong>the</strong>r, ate all those exotic food,<br />

and spaced out on <strong>the</strong> cola breaks. Let’s make our<br />

friendship a decade! Dear Atilla, my <strong>VU</strong> gym partner,<br />

you know that I was so strict on you during <strong>the</strong> yoga<br />

class or at <strong>the</strong> gym, because <strong>of</strong> “good intention and<br />

tough love”, right? And now stop calling me B****!!!<br />

Dennis Verzijl and Henry Vischer, thank you for<br />

your helps and comments through <strong>the</strong>se years. Dear<br />

Dennis, special thanks to <strong>the</strong> birthday present you<br />

gave me last year. It was ei<strong>the</strong>r “Life-threatening” or<br />

“Life-changing” experience. Fortunately, it is <strong>the</strong> latter<br />

one in my case. I cannot wait for more goodies. Dear<br />

Henry, aka Nemo, I love it when you sometimes<br />

dropped by my room (N-249) for a quick chat and also<br />

many thanks to your comments when I had to prepare<br />

my presentations. I hope your next holiday destination<br />

will be Thailand again. Besides <strong>the</strong> temple full <strong>of</strong><br />

tigers, we also have monkeys and o<strong>the</strong>r exotic animals,<br />

you know!<br />

Saskia Hulscher and Obbe Zuiderveld, thank you for<br />

your numerous helps. Dear Saskia, besides helping me<br />

with all <strong>the</strong> ordering <strong>of</strong> expensive kits that I needed,<br />

you also provided me something with much greater<br />

value in all <strong>the</strong>se years; “a friendly smile, a hello every<br />

time we met and <strong>the</strong> warm friendship”. Dear Obbe,<br />

thank you for trying to treat me very gently, although I<br />

sometimes make a mess in <strong>the</strong> RNC. As a sign <strong>of</strong> my<br />

appreciation, I will make sure to serve good beverages<br />

at my reception.<br />

My special warm gratitude also go to my former<br />

bench-mates and now FB friends; David Maussang,<br />

Leontien Bosch and Erik Slinger. I really enjoyed<br />

your presence and company. You made coming to <strong>the</strong><br />

149 | P a g e<br />

lab much more fun and time did go by quickly. Dear<br />

David, since this is possibly <strong>the</strong> only book I am going<br />

to write, I will come out <strong>of</strong> my closet now; you are my<br />

number one “lieveling” in <strong>the</strong> lab. Dear Leo and Erik,<br />

thank you for your supports and positive energy. I<br />

know I can always count on you, especially when it<br />

comes to moving. Thank you!<br />

My “Die-hard” yoga team; Anne, Azra, Chimed, Kim,<br />

Ellen, Meritxel, Marija, Petra, Saskia (N) and<br />

Sabrina. Thank all <strong>of</strong> you for <strong>the</strong> lovely and vibrant<br />

yoga evenings we shared toge<strong>the</strong>r at <strong>the</strong> <strong>VU</strong> library,<br />

<strong>the</strong> c<strong>of</strong>fee room and my apartment. Chest up! Stomach<br />

in! butt tights! Straight! and strong! Now ano<strong>the</strong>r 10<br />

sun salutations! No need to say more, girls. Your<br />

enthusiasms really made it happened!<br />

My students; T<strong>of</strong>an, Laura and Kim, I hope you did<br />

learn something about protein purification…besides<br />

scraping million <strong>of</strong> dishes. Thank you for all <strong>the</strong> fun,<br />

good laughs and youthful energy you bring along,<br />

especially Kim. Lieve Kim, thank you for being my<br />

personal assistant, my secretary, my dear friend, my<br />

sister and most importantly thank you for teaching me<br />

that I would never be <strong>the</strong> mature one in our<br />

friendship, but that it was also fine because you totally<br />

had it covered!<br />

I also wish to express my warm regards to all <strong>the</strong> o<strong>the</strong>r<br />

colleagues and <strong>the</strong> interns that I have encountered<br />

through <strong>the</strong>se years; Regina (thank you for <strong>the</strong><br />

friendship, I hope to visit you one day in Ghana), Jan<br />

Paul Bebelman (thanks for all <strong>the</strong> helps and many<br />

nice conversations we have), Jane van Heteren (I did<br />

my first 7 hills running because <strong>of</strong> you), Fiona<br />

Shenton (although only 2-3 days, but you taught me<br />

so many tips for <strong>the</strong> running), Mirabella (my all-time<br />

beautiful girl), Martine (MvK, thanks for playing<br />

tennis with me), <strong>the</strong> twin sisters (Angelique and<br />

Payman) and <strong>the</strong> twin bro<strong>the</strong>rs (Lars and Deni) who<br />

always put smiles on my face and many more <strong>of</strong> you<br />

that I cannot mention all in this chapter. Thank you!<br />

The “thuisfront”<br />

Dear Anneli Lundgren, you are one <strong>of</strong> <strong>the</strong> most<br />

kindhearted people I know-genuinely. Even though,


we are far apart for many years already, somehow I<br />

feel as if <strong>the</strong> distance just enriches our friendship.<br />

Thank you for all <strong>the</strong> postcards you sent to me from all<br />

over <strong>the</strong> world. I can hardly wait to see your next<br />

adventure. Dear Maria Edman, thank you for showing<br />

me how to be best friend to somebody. Dear Yanchao<br />

Huang and Mika, thank you for always trying to keep<br />

in touch. I hope one day our path will be crossed again.<br />

Dear Molly Ackerman, thank you for being a good<br />

role model for me, you are my artist, my inspiration<br />

and my special friend that I always look up to. My girls<br />

Siew Hoon Chan, Lianne Begga, and Julia<br />

Schorestene, you probably do not know each o<strong>the</strong>r,<br />

but you all shared very special characters that I<br />

appreciate so much in a friend. Thank you for always<br />

be <strong>the</strong>re for me when I am down and for showing me<br />

how cool it is to be independent!<br />

Many thanks also go to my Thai contacts:<br />

entire extended family members for providing a loving<br />

home for me. Without <strong>the</strong>ir love, supports, and<br />

ขอบคุณพี่หน�อย พี่ต�าย พี่เอ�ม อาจารย�ตั้ม encouragements อาจารย�แดง it would have been impossible for me<br />

เจ�าแม็คน�องสุดที่รักของพี่และครอบครัวของแม็คพี่เล็ก<br />

to get through <strong>the</strong> decade <strong>of</strong> cold and lonely time in<br />

และพี่ๆที่ร�านสบายๆและร�านศรีโสธรนะคะ Europe.<br />

ขอบคุณสําหรับอาหารอร�อยๆและไออุ�นที่มีให�จิ๊บเสมอมา<br />

พี่โจพี่โบเปรียบเสมือนคุณพ�อคุณแม�คนที่สองขอ<br />

ขอบคุณที่รักและดูแลหนูตลอดมา<br />

Special thanks to two <strong>of</strong> my paranimfen; Sonia Faut,<br />

my swee<strong>the</strong>art, I wish to thank you for <strong>the</strong> very<br />

intense six months we had toge<strong>the</strong>r in Basel. I<br />

genuinely feel as though I have found a soul sister in<br />

you. Thank you for being my friend, for showing me<br />

that it takes 8 doors between 1978 and 1985, for<br />

always checking if I am working on <strong>the</strong> <strong>the</strong>sis or not,<br />

for making <strong>the</strong> original French pancake for me, for<br />

learning me how to pronounce Mulhouse correctly, for<br />

suffering in <strong>the</strong> gym and firmenlauf with me, for<br />

hugging me when I cry and for making time for my<br />

promotion. “Sonia, ma chérie, Tu me fais sourire. Je<br />

t'aime de tout mon coeur.” Alejandro Bernadis, Dear<br />

Alejandro, you are probably <strong>the</strong> happiest man on earth<br />

and I always have a deep respect on you. I feel warm,<br />

comfortable and calm around your presence (which is<br />

probably what I need <strong>the</strong> most before <strong>the</strong> defense). It<br />

is my honor that you agreed to wear <strong>the</strong> penguin suit<br />

and be my paranimf. Muchísimas gracias!<br />

A very special thank is due to <strong>the</strong> Bikram yoga school<br />

<strong>of</strong> <strong>Amsterdam</strong>. It started out as “Jib’s new challenge”,<br />

but <strong>the</strong> integrity and <strong>the</strong> devotion <strong>of</strong> <strong>the</strong> fellow yogis<br />

150 | P a g e<br />

towards <strong>the</strong> practice deeply inspired me and<br />

completely changed my attitude and lifestyles. Also<br />

from <strong>the</strong>re, I got a chance to get to know many people<br />

outside <strong>the</strong> scientific world and even more<br />

importantly some <strong>of</strong> <strong>the</strong>m have become <strong>the</strong> dearest <strong>of</strong><br />

my heart.<br />

Dear Rodrigo, thank you for beating my ego down to<br />

nothing, so I could rise up completely as <strong>the</strong> true<br />

person. Thank you for taking risks with me, for<br />

embracing our differences, for continually surprising<br />

me with how wonderful your ideas are, for making me<br />

realize that having those huge dreams is more than<br />

okay, for giving your everything to everything and<br />

most <strong>of</strong> all, thank you for being “you”.<br />

I wish to thank my sister พี่โจ<br />

พี่โจ , my bro<strong>the</strong>r พี่โบ พี่โบ<br />

พี่โบ and my<br />

หนูขอให�เรามีกันและกันตลอดไป<br />

Lastly and most importantly, I wish to express my<br />

deepest gratitude to my parents, Mongkhol and<br />

Nonthawan Sansuk. You brought me into this<br />

world, raised me with all your love and care,<br />

supported me to your utmost ability, taught me to<br />

be a good person, and loved me unconditionally.<br />

For you, a thousand times over.<br />

To you, I dedicate this <strong>the</strong>sis.<br />

คุณพ�อคุณแม�คะ<br />

หนูทําสําเร็จแล�วนะคะความฝ�นของเรา<br />

ต�อไปนี้เราจะไม�ต�องอยู�ไกลกันแล�วนะคะ<br />

หนูรักคุณพ�อคุณแม�มากที่สุดในโลก


“The foot feels <strong>the</strong> foot when it feels <strong>the</strong> ground.”<br />

Lord Buddha

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!