11.03.2014 Views

Thesis final - after defense-7 - Jacobs University

Thesis final - after defense-7 - Jacobs University

Thesis final - after defense-7 - Jacobs University

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

A proteome wide approach to understand<br />

chromatography behavior on hydrophobic<br />

adsorbents<br />

by<br />

Nawab Ali<br />

A <strong>Thesis</strong> submitted in partial fulfillment<br />

of the requirements for the degree of<br />

Doctor of Philosophy<br />

in Biochemical Engineering<br />

Approved Dissertation Committee<br />

__________________________________<br />

Prof. Dr. Marcelo Fernández-Lahore<br />

Prof. Dr. Matthias Ullrich<br />

Dr. Abid Riaz Ahmad<br />

Date of Defense:<br />

June 05 th , 2012<br />

School of Engineering and Science


Acknowledgement<br />

All praise for Almighty ALLAH, Without Allah’s help, I would not be able to achieve<br />

anything in my life.<br />

I am greatly honored to owe my deepest thanks to my thesis supervisor Prof. Dr. Marcelo<br />

Fernández-Lahore for giving me the opportunity to carry out my PhD work under his kind<br />

supervision and accepting me in his international group. I greatly appreciate his valuable<br />

suggestions and supervision in my PhD work. I am highly thankful to him for accepting my<br />

wife as a graduate student under his kind supervision. He has given me an opportunity to<br />

perform my tasks independently which increased my curiosity for research and which<br />

accomplishes the true spirits of PhD. I am thankful for his student friendly and cooperative<br />

attitude in the last four years.<br />

I am also thankful to Prof. Dr. Matthias Ullrich and Dr. Abid Riaz Ahmad for their comments<br />

and nice suggestions for the improvement of the thesis.<br />

I also want to thank and acknowledge Muhammed Mobashir and Abdul Wakeel for helping<br />

me in the bioinformatics part of this work.<br />

I convey my heartiest acknowledgements to my group fellows Aasim, Zia, Nadia, Noor Shad,<br />

Farhat, Amir, Tuhidul Islam, Sirma, Amira, Rajesh, Dr Rami, Dr. Sonja, Miss. Nina, Antonio,<br />

Doreen, Iza, Naveen, Prasad, Rodrigo and Zumrad.<br />

Words are lacking to present my feeling for the prayers and struggles done by my parents<br />

throughout my life from childhood until now. I am now in this position due to my father<br />

continuous guidance and efforts during my education. I am also thankful to my brother<br />

Naveed Ali, sisters and brother in law; Sadeeq and Sabir for their love and care at each step of<br />

my life. I also want to acknowledge my other relatives and uncles Mumtaz Ali, Ihsan Ali,<br />

Qamer Zaman and my cousins Nazrali, Mohsin, Jenab, Aasim, Zafar, Amin, Khaild.<br />

I


I also want to mention the efforts of my wife Nadia for her patience and support during my<br />

PhD studies when I was in tough time of having funding and other research problems. I am<br />

also thankful to her for letting me free from other homely responsibilities when I was busy in<br />

my research work.<br />

I also want to convey my heartiest and sincerest acknowledgements to my closest friends<br />

Ammar, Qasim, Aasim, Yaqoob, Noor Muhammad saib, Abunaser saib, Farhan, Hamish saib,<br />

Farrukh saib, Roohalamin saib, Noor Shad, Zia, Wakeel, Sadiq, Amna, Farhat, Qazi, Hazir,<br />

Tariq, Inam, Tahir, Ali, Imran, Rehan, Saleem, Shahzad.<br />

Finally, I wish to thank Kohat <strong>University</strong> of Science and Technology (KUST), Khyber<br />

Pukhtunkhwa, Pakistan and Higher Education Commission, Islamabad, Pakistan for providing<br />

me opportunity for higher studies abroad.<br />

II


Nomenclature<br />

S. cerevisiae Saccharomyces cerevisiae<br />

HIC<br />

Hydrophobic interaction chromatography<br />

ASH<br />

Average surface hydrophobicity<br />

PDB<br />

Protein databank<br />

AH<br />

Average hydrophobicity<br />

ASA<br />

Accessible surface area<br />

CW<br />

Cowan-Whittaker scale<br />

MJ<br />

Miyazawa-Jernigan scale<br />

T<br />

Tanford scale<br />

AF<br />

Average flexibility<br />

CV<br />

Column volume<br />

mAU<br />

Milli absorbance unit<br />

IEF<br />

Isoelectric focussing<br />

TEMED<br />

N, N, N, N-tetramethyl-ethylenediamine<br />

DTT<br />

Dithiothreitol<br />

Mili Q water water purified with millipore system<br />

SDS-PAGE Sodium dodecyl polyacrylamide gel electrophoresis<br />

MALDI-ToF-MS Matrix Assisted Laser Desorption/Ionization Time-of-Flight Mass spectrometry<br />

2-D PAGE Two dimensional polyacrylamide gel electrophoresis<br />

PS<br />

Phenyl-Sepharose<br />

SP<br />

Source-Phenyl<br />

TP<br />

Toyopearl-Phenyl<br />

TB<br />

Toyopearl-Butyl<br />

TH<br />

Toyopearl-Hexyl<br />

TE<br />

Toyopearl-Ether<br />

S j<br />

S i<br />

Sum of total surface area of the proteins<br />

Sum of accessible surface area of for all the amino acids of type i<br />

r i Superficial area of amino acid for all the amino acids of class i.<br />

φ<br />

i<br />

Hydrophobicity of amino acids for all the amino acids of class i<br />

Φ surface Average surface hydrophobicity<br />

Mr<br />

pI<br />

Molecular weight of the protein<br />

Isoelectric point of the protein<br />

III


Dedication<br />

To my father Haji Sher Afzal and mother Kamkhwaba for their continuous support and<br />

efforts during my education and in every aspect of my life.<br />

IV


Abstract<br />

The separation behavior of the Saccharomyces cerevisiae cell proteome was explored by<br />

hydrophobic interaction chromatography (HIC). Several beaded adsorbents were studied. A<br />

first group of HIC adsorbents (n = 3) harbored the same ligand moiety (i.e., Phenyl) but<br />

presented differences in the chemical nature of the matrix backbone. On the other hand, a<br />

second group of adsorbents (n = 3) presented the same chemical structure (i.e. based on<br />

polymethacrylate; Toyopearl / TP) but differed in the chemical nature of the HIC ligands.<br />

Chromatography runs were performed under typical conditions employing ammonium sulfate<br />

/ phosphate buffer (pH = 7.5) as a mobile phase. Chromatography fractions were collected<br />

and further analyzed by two-dimensional polyacrylamide gel electrophoresis (2-D PAGE);<br />

selected spots were identified by MALDI-ToF-MS and database search.<br />

A comparative analysis of protein separation behavior is presented as a function of the<br />

adsorbent type. As expected, TP-Hexyl and TP-Butyl showed increased protein retention in<br />

comparison with TP-Ether. Moreover, an influence of protein size and isoelectric point (pI)<br />

was also revealed. Larger and / or neutral proteins showed increased while acidic (pI < 6) and<br />

basic (pI > 8) proteins depicted decreased or increased retention, depending on the adsorbenttype.<br />

Protein characteristics such as average surface hydrophobicity, average hydrophobicity<br />

average polarity, average bulkiness and average flexibility were tabulated for the identified<br />

spots and correlated with chromatography behavior. The proteins properties revealed limited<br />

ability to differentiate among the ligands and base supports for their hydrophobicity. This<br />

work -for the first time- studied the separation behavior of a real and complex cell proteome<br />

as a function of commercially available HIC adsorbents. The mentioned results will favor the<br />

understanding and the application of HIC, a method of choice in many industrial bioseparation<br />

schemes.<br />

I


Table of Contents<br />

1. Introduction ............................................................................................................................ 1<br />

1.1. Introduction to Biotechnology ........................................................................................ 1<br />

1.2. Downstream processing .................................................................................................. 2<br />

1.3. Hydrophobic interaction chromatography (HIC) ............................................................ 4<br />

1.4. Factors affecting protein chromatography behavior in HIC ........................................... 7<br />

1.4.1. Effect of mobile phase parameters ........................................................................... 7<br />

1.4.2. Effect of stationary phases in HIC ......................................................................... 10<br />

1.4.3. Effect of protein properties in HIC ........................................................................ 12<br />

1.4.4. The comparison between HIC and RPC ................................................................ 13<br />

1.5. Analysis of the yeast proteome as biotech feedstock .................................................... 15<br />

1.6. Mass spectrometry and database searching ................................................................... 16<br />

1.7. The in silico Approach .................................................................................................. 18<br />

1.8. Goal of the work ............................................................................................................ 21<br />

2. Experimental approach ......................................................................................................... 24<br />

2.1. Materials ........................................................................................................................ 24<br />

2.2. Experimental strategy .................................................................................................... 24<br />

2.2.1. Cell Cultivation ...................................................................................................... 26<br />

2.2.2. Preparation of the crude extract ............................................................................. 26<br />

2.2.3. Chromatographic fractionation .............................................................................. 26<br />

2.2.4. Membrane dialysis ................................................................................................. 27<br />

2.2.5. Two dimensional polyacrylamide gel electrophoresis (2-D PAGE) ...................... 27<br />

2.2.6. Protein content of the fractions .............................................................................. 29<br />

2.2.7. Protein identification by MALDI-ToF-MS ............................................................ 29<br />

2.2.8. Characterization of proteins utilizing Bioinformics ............................................... 31<br />

3. Results and Discussion ......................................................................................................... 35<br />

3.1. Influence of the different ligand chemistries during hydrophobic interaction<br />

chromatography ........................................................................................................................ 35<br />

3.1.1. Summary .................................................................................................................... 35<br />

3.1.2. Scientific background of the approach ....................................................................... 36<br />

3.1.2.1. Hydrophobic adsorbents ...................................................................................... 36<br />

3.1.2.2. Protein properties in the context of HIC ............................................................. 39<br />

3.1.2.3. Towards an in silico approach for the downstream processing of bioproducts .. 43<br />

2


3.1.3. Results and Discussion ............................................................................................... 45<br />

3.1.3.1. Chromatographic profiles of different ligands .................................................... 45<br />

3.1.3.2. Two dimensional gel electrophoresis and proteins identification ....................... 47<br />

3.1.3.3. Influence of the ligands chemistries and average protein properties .................. 55<br />

3.1.3.3.1. Average hydrophobicity (AH) ............................................................................. 55<br />

3.1.3.3.2. Average polarity (AP) .......................................................................................... 59<br />

3.1.3.3.3. Average flexibility (AF)....................................................................................... 62<br />

3.1.3.3.4. Average Bulkiness (AB) ...................................................................................... 65<br />

3.1.3.3.5. Average surface hydrophobicity (ASH)............................................................... 66<br />

3.1.3.4. Comparative studies of the hydrophobic ligands based on average protein<br />

properties .......................................................................................................................... 68<br />

3.1.3.5. Application note .................................................................................................. 70<br />

3.1.4. Partial conclusions ...................................................................................................... 71<br />

3.2. Influence of the different base support chemistries during hydrophobic interaction<br />

chromatography ........................................................................................................................ 72<br />

3.2.1. Summary .................................................................................................................... 72<br />

3.2.2. Chemistry of the base supports .................................................................................. 73<br />

3.2.3. Results and Discussion ............................................................................................... 75<br />

3.2.3.1. Chromatographic fractionation ........................................................................... 75<br />

3.2.3.2. The electrophoretic separation and proteins identification ................................. 78<br />

3.2.3.3. Influence of the base support chemistries and average protein properties .......... 85<br />

3.2.3.3.1. Average hydrophobicity (AH) ............................................................................. 85<br />

3.2.3.3.2. Average polarity (AP) .......................................................................................... 89<br />

3.2.3.3.3. Average surface hydrophobicity (ASH)............................................................... 93<br />

3.2.3.3.4. Other protein properties affecting the chromatographic behavior ....................... 95<br />

3.2.3.4. Comparative studies of the different base supports based on average protein<br />

properties .......................................................................................................................... 98<br />

3.2.3.5. Application note ................................................................................................ 100<br />

3.2.3. Partial conclusions .................................................................................................... 101<br />

3.3. Protein distributions on 2-D gels: the influence of molecular weight and isoelectric point<br />

coordinates ............................................................................................................................. 102<br />

3.3.1. Summary .................................................................................................................. 102<br />

3.3.2. Results and Discussion ............................................................................................. 103<br />

3.3.2.1. Unfolding of proteins during HIC ..................................................................... 103<br />

3


3.3.2.2. Protein distributions on 2-D gels ....................................................................... 104<br />

3.3.2.3. The effect of protein molecular mass on the chromatographic behavior .......... 106<br />

3.3.2.3. The influence of protein isoelectric point on the chromatographic behavior ... 108<br />

3.3.3. Partial conclusions .................................................................................................... 111<br />

4. Conclusions and Remarks .................................................................................................. 113<br />

4.1. General conclusions and remarks ................................................................................ 113<br />

4.2. Influence of different ligands ...................................................................................... 113<br />

4.3. Influence of different base supports ............................................................................ 113<br />

4.4. Effect of protein properties on separation behavior during chromatography ............. 114<br />

4.5. 2-D gels analysis ......................................................................................................... 115<br />

4.5. Additional conclusions ................................................................................................ 115<br />

4.6. Recommendations for future work .............................................................................. 116<br />

5. References .......................................................................................................................... 117<br />

4


Chapter 1<br />

INTRODUCTION<br />

5


Chapter 1<br />

1. Introduction<br />

1.1. Introduction to Biotechnology<br />

The term biotechnology is a fusion of biology and technology which is concerned with the<br />

utilization of biological agents to produce useful products of human use (1). Biotechnology<br />

has been known for thousands of years, where humans have used it for the food purposes e.g.<br />

for the selective breeding of crops and livestock, production of wine, beer, cheese and other<br />

dairy products. In the early twentieth century, a new era of the modern biotechnology started<br />

with the production of complex organic molecules like antibiotics, vaccines and monoclonal<br />

antibodies. The production of these macromolecules has been optimized by the improved<br />

fermentation procedures and novel downstream processing approaches (2). The discoveries<br />

for the new products have made the pharmaceutical industry as one of the fastest growing<br />

sectors in the global economy. The pharmaceutical, agrichemical and biotechnology related<br />

products have the billions dollar annual sale in market excluding the food and beverages (3, 4).<br />

Proteins and enzymes are the main biopharmaceutical products and have several applications<br />

in food and medical biotechnology (5-7). The production of human biopharmaceuticals was in<br />

limited supply and very costly before the recombinant DNA technology revolution. The<br />

proteins were commonly used before from the plant's and the animal's origin directly which<br />

was mostly not feasible due to the resulting immune responses against foreign antigens.<br />

However, genetic engineering made it possible to clone the responsible genes in microbial<br />

cells and produce the larger amount of protein of interest. Bioprocess can be usually divided<br />

into two main parts such as upstream and downstream processing. Upstream processing is the<br />

successful fermentation step and downstream processing is the purification plan of the desired<br />

protein at a large scale. Although a lot of work has been done for the fermentation<br />

optimization and development of recombinant strain producing the biomolecules, but the<br />

purification process is still a major bottleneck in order to fulfil the market demands. The main<br />

1


Chapter 1<br />

objective is to obtain the highest yield and purity of the target protein in the minimum<br />

possible resources by downstream processing (8).<br />

1.2. Downstream processing<br />

Purification is the next step <strong>after</strong> the production of a biomolecule through successful<br />

fermentation. The isolation and purification of a biomolecule at industrial level to a form,<br />

suitable for its intended use is called downstream processing. The products of biotechnology<br />

include whole cells, organic acids, amino acids, antibiotics, industrial enzymes, gums,<br />

vaccines etc. The different separation methods can be applied for the recovery and<br />

purification of the biomolecules depending on their sizes and natures (2). The purification of<br />

the desired product needs certain downstream processing steps, which accomplishes 60% of<br />

the total cost, excluding the cost to purchase raw materials (9, 10). Several expert systems<br />

have been developed for the protein purification processes based on the use of empirical<br />

knowledge that is not exercised in nature and is typically used by experts (11-14). The<br />

number of steps varies ranging from eight to fourteen in an expert system of common practice.<br />

Additional steps will cause an increase in the cost and also a decrease in the yield. As the<br />

number of steps for any downstream process is increasing, the product loss will occur. For an<br />

efficient downstream processing, it is important to design a bioprocess with the minimum<br />

possible steps. This will increase the yield and purity of the product to an acceptable range<br />

(15-18). The bioprocess has usually two main steps; recovery and purification. Recovery is<br />

composed of a collection of cell proteome from disrupted cells, while purification is the<br />

chromatographic fractionation of the cell proteome for the isolation of desired protein. The<br />

target protein has to be isolated from other impurities such as DNA, RNA and cellular debris,<br />

if the product exists intracellular. Minute amount of other impurities can be separated by<br />

different chromatographies to avoid any toxic effects in the <strong>final</strong> product coming from a<br />

biological mixture (8). It is also important to study other contaminants available in the host<br />

2


Chapter 1<br />

cell proteome to facilitate the separation of the target protein in the real bioprocess. Several<br />

expert systems have been proposed to link the efficiency of a unit operation with the protein<br />

properties and to separate the target protein from the other discrete number of contaminants.<br />

Most of these approaches to understand the chromatographic behavior of proteins were<br />

limited to the use of model proteins instead of protein mixtures in the host cell proteome. No<br />

comprehensive struggle has been made before to define the contaminant proteome of real<br />

recombinant host in order to investigate the individual contaminant behavior during its<br />

chromatographic separation. A paradigm shift from focussing on protein product to defining<br />

the major contaminant might give a more realistic picture and could also make this approach<br />

more suitable. The characterization of host related proteins is also important to maintain the<br />

product quality (9, 19).<br />

Chromatography is an important unit operation of any downstream process which helps to<br />

remove the contaminants during the purification of biological materials. Chromatographic<br />

separation of protein mixtures is becoming one of the most effective and widely used means<br />

of purifying individual proteins. The chromatographic methods for protein separations are<br />

based on the general physicochemical properties of the proteins (Figure 1) (20). Minor<br />

differences between various proteins such as size, charge and hydrophobicity can be used to<br />

purify one protein from the other proteins during chromatography. Chromatographic<br />

separations are usually based on protein characteristics such as charge (ion exchange<br />

chromatography), size (size exclusion chromatography), isoelectric point (chromato focussing)<br />

affinity (affinity chromatography) and hydrophobicity (hydrophobic interaction<br />

chromatography and reverse phase chromatography). Hydrophobic interaction<br />

chromatography (HIC) is a chromatographic method based on protein hydrophobicity (4).<br />

HIC was considered as more significant than other chromatographic methods, due to its mild<br />

hydrophobic nature and feasibility for the separation of less stable proteins (21). HIC is a<br />

3


Chapter 1<br />

chromatographic method of choice in downstream processing and have been widely used for<br />

the purification of proteins, enzymes, antibiotics and vaccines (14, 22, 29).<br />

Figure 1: The chromatographic methods based on different protein properties (Adapted from<br />

Ref. 20).<br />

1.3. Hydrophobic interaction chromatography (HIC)<br />

In industrial biotechnology, the purification of the desired product is very important <strong>after</strong><br />

successful fermentation. There are several chromatography methods available for protein<br />

purification based on the physicochemical properties of the proteins. However, HIC can be<br />

preferred due to its ability of rapid separation, high resolution and less chances of protein<br />

denaturation (23, 24). HIC is based on the reversible interactions between the hydrophobic<br />

surface patch of a protein and the hydrophobic surface of a chromatographic medium at<br />

moderately high salt concentrations. Proteins contain a variety of amino acids having<br />

hydrophilic and hydrophobic residues. Many of the hydrophobic residues are buried in the<br />

interior of the proteins, but on exposure to a solvent, a significant proportion of them can<br />

appear on the surface. However, the ratio of surface hydrophobicity differs in different<br />

4


Chapter 1<br />

proteins. It is this difference which can be exploited in HIC for protein separations. Protein<br />

hydrophobicity is the main physicochemical property which has a decisive role in retention<br />

during chromatography. There is no universally agreed single method to calculate protein<br />

hydrophobicity, however there is a general agreement that it can be determined by the<br />

hydrophobic contribution of the amino acid residues (25, 26). The operating conditions such<br />

as mobile phase properties (salt type, ionic strength and pH), stationary phase characteristics<br />

(chemical nature of the backbone, type of hydrophobic ligand and substitution level of the<br />

resin) and temperature play important roles in HIC (27, 28). Hydrophobic interactions are the<br />

most important non covalent forces which have the ability to maintain the native structures of<br />

the proteins at high salts concentrations. The basic principle of the HIC is different from size<br />

exclusion and ion exchange chromatography (IEC) and thus can be used for the products<br />

which have no possibility to be separated by these techniques. HIC is preferred over other<br />

chromatographic methods if proteins of similar size or isoelectric point coordinates have to be<br />

separated from each other. HIC can also be used as an ideal step, <strong>after</strong> the separation of<br />

biomaterials during IEC at high salt concentrations. Due to the entries of several new HIC<br />

media and better understanding of the factors affecting the retention, HIC has become one of<br />

the most widely used method for purification of proteins such as serum proteins, membrane<br />

proteins, nuclear proteins, receptors and recombinant proteins for research and industrial<br />

applications (14, 23, 29). The enzymes of commercial importance such as lipases, amylases,<br />

streptokinases can be purified utilizing HIC (30). Several expert systems have been produced<br />

to exploit the physicochemical properties of the proteins in order to separate the desired<br />

product from other contaminants in HIC (31, 32). In most of these approaches, the prediction<br />

of protein retention was limited to the use of model proteins instead of the host cell proteome<br />

(33). Studies of the host cell proteome which produces the product of interest can help in the<br />

design of the process; give in depth knowledge of the target and undesired proteins for the<br />

efficient downstream processing. No large scale effort has been made to study the<br />

5


Chapter 1<br />

contaminants which have close similarity to the target protein in the host cell proteome (19).<br />

In order to do so, one has to design the experiments to analyze the adsorption affinity of the<br />

adsorbents with a natural host.<br />

The first report about HIC was done by Sheppard and Tiselius observing the retention of dye<br />

in the presence of sulfate and phosphate solutions and was called salting out chromatography<br />

(34). Afterwards, Shalteil and Er-el used the term hydrophobic chromatography, Hofstee<br />

named the method as hydrophobic adsorption chromatography (34, 35). Finally, Hjerten in<br />

1973 described the method as HIC (36). Porath et al. discovered that hydrophobic adsorption<br />

was enhanced by using salts like sodium chloride or phosphate chloride and named the<br />

method as salt promoted adsorption chromatography (34).<br />

A typical chromatogram obtained during my experiments has been presented in Figure 2. To<br />

facilitate the hydrophobic interactions and adsorption of proteins, it is important to load the<br />

proteins on a column with a buffer of high salt concentrations, followed by elution with a<br />

linear or a stepwise decrease in the salt concentration (37). Linear gradient can be applied<br />

during chromatographic separation of biological mixtures to separate the proteins of diverse<br />

nature. After sample application, the proteins with no binding affinity will elute in flow<br />

through and washing step. The proteins bound with the adsorbent will elute according to their<br />

binding capabilities and the extent of protein hydrophobicity. The proteins with less surface<br />

hydrophobic residues will elute at the start of the chromatography. In contrast, the proteins<br />

with high hydrophobic residues on the surface will be tightly bound and will elute at the end<br />

of the chromatography. The highly hydrophobic proteins are usually tightly bound to the<br />

column and can be removed by salt free buffer or 20% ethanol. As the surface hydrophobicity<br />

of the proteins is increasing, the retention time of the proteins will increase accordingly<br />

during chromatography.<br />

6


Chapter 1<br />

Salt conc. decreasing<br />

(Gradient elution)<br />

Sample application.<br />

Abs.<br />

280 nm<br />

Unbound proteins<br />

elute before gradient.<br />

Tightly bound<br />

proteins elute in salt<br />

free conditions<br />

Column Volume (CV)<br />

Figure 2: A typical HIC chromatogram describing the chromatographic fractionation during HIC.<br />

1.4. Factors affecting protein chromatography behavior in HIC<br />

1.4.1. Effect of mobile phase parameters<br />

The mobile phase parameters affecting the protein retention are usually ionic strength, type of<br />

salts and buffer pH. Protein adsorption on hydrophobic adsorbents is usually favored by<br />

moderately high salt concentrations. The concentration of the salt should be according to the<br />

binding strength of the proteins and adsorbents and below the concentration causing<br />

precipitation of proteins. The optimum concentrations of salts are usually between 0.75 M and<br />

2.0 M with ammonium sulfate and 1.0 M to 4.0 M with sodium chloride (27). The different<br />

types of salts can be divided into kosmotrophic and chaotrophic salts depending on their<br />

ability of hydrophobic interactions. The chaotrophic salts (magnesium sulfate and magnesium<br />

chloride) have less capability to bind water molecules which increases the inclusion of water<br />

molecules on the protein and ligand surface and thus decreases hydrophobic interactions<br />

7


Chapter 1<br />

(salting in effect). While kosmotrophic salts (sodium, potassium or ammonium sulfate) bound<br />

tightly to water molecules which facilitate the exclusion of water molecules on the proteins<br />

and ligand surface and thus promotes hydrophobic interactions (salting out effect) (Figure 3)<br />

(20, 38).<br />

Figure 3: The mechanism of hydrophobic interaction chromatography (Adapted from Ref. 20)<br />

It has been reported by Hjerten that the main force which arises during the dispersion of water<br />

molecules is an increase in the entropy of the system. It has been stated that the hydrophobic<br />

ligand-protein interaction was thermodynamically favorable process (36, 39). The Hofmeister<br />

(lyotrophic) series of cations and anions was proposed before, starting with those favoring the<br />

interactions to those that disfavor the hydrophobic interactions. For anions and cations, the<br />

series is given below.<br />

Anions; PO 4<br />

3-<br />

> SO 4<br />

2-<br />

> CH 3 COO - > Cl - > Br - > NO 3<br />

-<br />

> ClO 4<br />

-<br />

> I - > SCN -<br />

Cations; NH 4<br />

+<br />

> Rb + > K + > Na + > Li + > Mg 2+ > Ca 2+ > Ba2 +<br />

8


Chapter 1<br />

The ions at the start of the series are called kosmotropes or anti-chaotropes and have been<br />

reported to have stronger interactions with water molecules and promote hydrophobic<br />

interactions (40, 41). The effects of the salts on protein retention can also be explained by the<br />

number of water molecules released by the induction of different types of salts. The<br />

selectivity of the different salts can also be predicted by the differences in their capability to<br />

exclude water molecules from proteins and adsorbent surface (27).<br />

The mobile phase pH is another factor affecting the protein adsorption during hydrophobic<br />

interaction chromatography. At highly basic pH values (up to 9-10), a decrease in<br />

hydrophobic interactions between proteins and adsorbent occur, due to the changes in the<br />

hydrophilicity of proteins. In contrast, hydrophobic interactions increase by decreasing the pH<br />

of the mobile phase. The proteins which are usually not able to bind at neutral pH will have<br />

more chances of binding at acidic pH values (42). The basic proteins, such as lysozyme (pI =<br />

10.7) has shown longer retention at basic pH values. In contrast, acidic proteins, such as<br />

human serum albumin (pI = 5.2) has revealed less retention during chromatography at basic<br />

pH values (43). In another report, the total number of releasing water molecules were found<br />

higher, when the mobile phase pH was close to the pI of the protein and decreased when pH<br />

was away from the pI (44). The effect of pH has been rarely reported and the reason could be<br />

the instability of the proteins and adsorbents at elevated pH conditions. Another study also<br />

stated that an optimal pH should be maintained in HIC to obtain high purity and yield at their<br />

maximum extent (45-48). The temperature was also reported to have an effect in HIC. It has<br />

been observed that increasing the temperature promoted protein retention and lowering the<br />

temperature enhanced protein elution. At higher temperature, the proteins are usually<br />

denatured and hydrophobic residues come on surface which resulted in high hydrophobic<br />

interactions. Due to the higher chances of unfolding at elevated temperatures, the unstable<br />

proteins should be separated at low temperatures. The proteins purified in the cold room may<br />

not be reproducible at the room temperature (49, 50). All the reports for the effects of these<br />

9


Chapter 1<br />

factors were mostly studied with model proteins; however their applicability to a natural host<br />

still needs to be explored.<br />

1.4.2. Effect of stationary phases in HIC<br />

Stationary phases are usually composed of small hydrophobic groups (e.g. butyl, phenyl and<br />

hexyl) attached to the hydrophilic polymer base supports. The different types of stationary<br />

phases in HIC differ in the degree of ligand substitutions on the base support. The base<br />

supports can also differ in the chemical nature and particle size (51). The availability of the<br />

hydrophobic adsorbents is increasing in the market, due to the increasing applications of HIC<br />

as a purification method. The adsorbents determine the protein adsorption selectivity in HIC<br />

according to the nature of ligands and base supports. The straight chain alkyl ligands were<br />

reported for pure hydrophobic character while aryl ligands (phenyl) showed mixed mode<br />

behavior due to the presence of aromatic interactions in addition to the hydrophobic<br />

interactions (52-55). The polyether’s ligands such as polyethylene glycols and polypropylene<br />

glycols were also reported for their milder hydrophobic nature than alkyl ligands (56). The<br />

retention of cytochrome c, ribonuclease A, lysozyme and ovalbumin has been reported on<br />

polyvinyl alcohol, oligoamino alcohol, and polyoxyethylene silica supports in descending<br />

gradient of ammonium sulfate buffer. It has been observed that hydrophobic interactions<br />

between proteins and adsorbents increased with increasing chain lengths of the ligand (55, 57).<br />

However, resolution of the chromatography decreased when ligands of the higher chain<br />

lengths have been used (51). The effects of adsorbent hydrophobicity on the conformation of<br />

proteins were also reported. The elution profiles of lactalbumin have shown different behavior,<br />

when separated by polyether stationary phases having methyl and ethyl terminating groups.<br />

The lactalbumin has given one and two peaks for the adsorbents with methyl and ethyl<br />

ligands, respectively. The first peak was the native one while the second peak was the<br />

denatured peak of the lactalbumin. The second peak has been retained longer due to the<br />

10


Chapter 1<br />

exposure of hydrophobic residues usually buried in the interior of proteins. Other protein such<br />

as lysozyme when applied has shown single peak on both adsorbents due to the stable nature<br />

of the protein (46). The charged type HIC adsorbents have shown different behavior during<br />

chromatographic separations. Ionic interactions will occur between proteins and adsorbents in<br />

the presence of charges, which favor the use of charge free adsorbents in HIC. The choice of<br />

ligand type is hard to interpret and could be established case to case basis depending on the<br />

biomaterials to be separated (53, 54). The retention behavior of the proteins can be changed<br />

by adjusting the ligand substitution on the base support. The protein binding capacity has<br />

been increased by an increase in the substitution degree of the ligand. Once the degree of<br />

substitution reaches beyond the sufficient level, the binding capacity of proteins remains<br />

constant, although the strength of interaction increases. At very high substitution of the<br />

ligands, multipoint attachments occur between proteins and adsorbents resulting in stronger<br />

binding. Thus harsh conditions are required to elute the product which may also increase the<br />

chances of protein denaturation. Some studies have preferred low ligand density in case of the<br />

unstable proteins to preserve the native structures of proteins (58-60).<br />

There are different types of base supports such as hydrophilic polysaccharide gels (e.g.,<br />

agarose and dextran), synthetic polymers (polymethacrylate and polystyrene divinyl benzene)<br />

or modified silica gel commonly used in HIC. It has been recently investigated with model<br />

proteins, that the adsorbent chemistry and up to some extent its particle size have effects on<br />

the retention behavior during chromatography (61, 62). The total number of water molecules<br />

released upon adsorption was found proportional to the hydrophobic area of the adsorbent<br />

(44). All the above findings were based on the retention data of model proteins and have been<br />

never investigated with the process proteomics approach.<br />

11


Chapter 1<br />

1.4.3. Effect of protein properties in HIC<br />

There are several protein properties affecting the chromatographic separation of proteins in<br />

HIC. However, the most important is protein hydrophobicity (51) and specifically average<br />

surface hydrophobicity (63) which has a decisive role in HIC. The protein hydrophobicity can<br />

be either the hydrophobicity of the exposed amino acid residues of a protein called as<br />

“average surface hydrophobicity” or depending on the hydrophobicity of the exposed and<br />

buried residues of a protein called as “degree of hydrophobicity” (27). The protein<br />

hydrophobicity has the main role in HIC which relies on no absolute definition. There is a<br />

general agreement that protein hydrophobicity can be the combined hydrophobic<br />

contributions of all the amino acids in a protein sequence. The hydrophobic potential of the<br />

various proteins can be derived by utilizing several amino acid hydrophobicity scales. Several<br />

experimental and theoretical approaches have been used by different researchers to assign the<br />

hydrophobicity value to each individual amino acid (25, 31, 64, 65). There are several<br />

hydrophobicity scales available in open literature which can be used for the determination of<br />

protein hydrophobicity. Tanford is one of those scales based on the transfer of free energies of<br />

the amino acid side chains from an organic environment to an aqueous environment (25, 66).<br />

Proteins when applied in an aqueous solution, the hydrophilic amino acid residues such as<br />

serine, threonine, asparagines, and glutamine tend to be exposed on the surface while most of<br />

the hydrophobic amino acid residues such as alanine, valine, leucine, isoleucine,<br />

phenylalanine, tryptophan, and tyrosine will be buried in the interior of protein molecules.<br />

However, adsorption of proteins occurs based on binding of the surface hydrophobic patches<br />

of a protein on the adsorbent. Several methods have been reported to quantify “protein surface<br />

hydrophobicity”. The first one was based on the binding of fluorescence probes to the<br />

hydrophobic region of a protein which has given rise to an increase in fluorescence according<br />

to protein hydrophobicity. Due to the possibility of ionic interactions in case of the anionic<br />

probes, this methodology was considered as non-applicable (64). The second method was<br />

12


Chapter 1<br />

based on the surface hydrophobic residues in a protein three dimensional structure (67). In<br />

addition to protein hydrophobicity, there are other physicochemical properties of the proteins<br />

such as average flexibility, average polarity, molecular weight and charge which have been<br />

rarely used for their effects in HIC. The average flexibility was reported to have a direct<br />

relationship with retention time of the model proteins in HIC. The proteins with low<br />

flexibility have been reported to be rigid towards conformational changes during<br />

chromatographic separations. In contrast, the protein with high flexibility will have more<br />

flexible nature towards conformational changes and when they come in contact of HIC<br />

surface, it will unfold or expose the internal surface in order to increase hydrophobic<br />

interactions and result in longer retention during chromatography (33, 62). The average<br />

polarity of the proteins was rarely reported to have an effect on protein retention in HIC. It<br />

was reported before that there is no obvious relationship between retention time and<br />

isoelectric point of the individual proteins (68). The effect of molecular mass on the retention<br />

behavior of proteins has been investigated before in ion exchange chromatography (19).<br />

However, there are very few reports where the effect of protein molecular weight on the<br />

chromatographic separation has been observed in HIC (33). There are few reports available<br />

where the above mentioned properties have been studied in HIC. However, these properties<br />

have been never researched for their effects on the chromatographic separations of a host cell<br />

proteome during HIC.<br />

1.4.4. The comparison between HIC and RPC<br />

The protein adsorption on hydrophobic adsorbents is generally weaker than reported in<br />

reverse phase chromatography (RPC), although both methods were based on hydrophobic<br />

interactions. In RPC, the protein-adsorbent interaction is so strong that protein elution can<br />

only be exerted by using organic solvents as elution buffer. The use of organic solvents and<br />

strong hydrophobic interactions are usually resulting in irreversible denaturation of proteins.<br />

13


Chapter 1<br />

This is also detrimental to the native structure and biological activity of proteins (69). In<br />

contrast, HIC is based on mild hydrophobic interactions and have importance to maintain the<br />

native structures and biological activity of the target proteins. The protein retention time can<br />

be predicted in HIC based on the surface residues of a protein. However, due to the<br />

denaturation of proteins in RPC, the elution is based on the exposed and buried hydrophobic<br />

residues of a protein (70). In other words, the protein retention in RPC can be estimated from<br />

the primary structure of a protein (71). In Table 1, the comparison between HIC and RPC has<br />

been given based on some general features (72) . There are several reports where they have<br />

used RPC for the purification of peptides. Several models have been proposed for the<br />

prediction of peptide retention time (73, 74). Although RPC has been widely used in the<br />

purification of peptides, however the application of RPC to large polypeptides and proteins<br />

has seen a significant increase in the last decade. The prediction of protein retention time has<br />

been reported based on the hydrophobicity and polypeptide chain length (75). However,<br />

further improvement to design prediction models is highly needed to be investigated.<br />

Table 1: The comparison of HIC and RPC (Adapted from Ref. 72).<br />

Description HIC RPC<br />

Stationary phases Low ligand density High ligand denisty<br />

Hydrophobic conditions Mild hydrophobic conditions High hydrophobic conditions<br />

Elution buffer Decreasing gradient of salts Increasing gradient of<br />

organic solvents<br />

Conformational changes Low conformational changes High conformational changes<br />

Protein retention<br />

Based on exposed<br />

Based on exposed and buried<br />

hydrophobic residues<br />

hydrophobic residues<br />

14


Chapter 1<br />

1.5. Analysis of the yeast proteome as biotech feedstock<br />

The proteomics tools such as two dimensional gel electrophoresis (2-D PAGE) and MALDI-<br />

ToF-MS are becoming the prominent methods for the analysis of the cell proteomes of<br />

various organisms. The proteomics analysis is usually the electrophoretic separation of<br />

proteins using 2-D PAGE followed by protein identifications (106). These proteomics studies<br />

are important to design the purification strategies and bioprocesses for the recombinant<br />

proteins. The bioprocesses are usually step by step approaches such as protein extraction,<br />

concentration, chromatographic separations and polishing. Several bioprocesses have been<br />

developed using different host cells and chromatographic methods for the purification of<br />

target proteins (109-110). Previously, master step purification was reported for the separation<br />

of Haemophilus influenzae cell proteome by heparin chromatography followed by<br />

electrophoretic separations and identification of proteins (76). The master step purification<br />

has gained more interest due to complete genome sequencing of several microorganisms of<br />

industrial interest and with the additional new improvements in the 2-D PAGE and MALDI-<br />

ToF-MS. The studies of the host cell proteome expressing the recombinant proteins called<br />

expression systems are important for protein purification. The expression systems can have<br />

the advantage of producing the maximum amount of target proteins in comparison to natural<br />

systems. There are few expression systems for the recombinant protein production such as<br />

bacteria, plants, animals, and yeast cells, usually Saccharomyces cerevisiae and Picchia<br />

pastoris (77). The cell proteomes of several microorganisms such as Candida albican,<br />

Plasmodium falciparum and Haemophilus influenzae has been reported by 2-D PAGE<br />

technology (78, 112). The Saccharomyces cerevisiae was selected as a host cell proteome due<br />

to several advantages over other expression systems such as high cell growth, high protein<br />

expression ratio, post translational modifications and most importantly it was the first<br />

organism to have its genome completely sequenced. The crude extract of the yeast cell<br />

15


Chapter 1<br />

proteome has been studied before using 2-D PAGE and mass spectrometry (105). However,<br />

the yeast cell proteome has been rarely investigated for its separation behavior on<br />

hydrophobic adsorbents by HIC. Hence, a study was planned to get a deeper insight into S.<br />

cerevisiae cell proteome and its separation behavior during HIC. The yeast cell proteome was<br />

chromatographed on different hydrophobic adsorbents and the fractions were collected to<br />

further investigate the proteins available in each chromatographic fraction. In a previous<br />

report, a three dimensional protein characterization approach has been applied to a complex<br />

plant extract using aqueous two phase system instead of HIC (112). The separation of yeast<br />

cell proteome was also three dimensional (3D) based on hydrophobicity (HIC), followed by<br />

its separation based on isoelectric point and molecular weight (2-D PAGE). Thus, the<br />

chromatographic and electrophoretic separation of the proteome was mainly based on three<br />

basic properties of the proteins, such as molecular weight, isoelectric point and<br />

hydrophobicity.<br />

1.6. Mass spectrometry and database searching<br />

Mass spectrometry is an important step in most of the proteome wide approaches to<br />

understand the biological processes. During the last decade, the genomes of most of the<br />

organisms of biological importance have been sequenced for the better understanding of<br />

genomes and proteomes. Once the genome sequence information was collected, the paradigm<br />

has shifted from sequencing to identification. The current paradigm was to investigate the<br />

host cell proteomes through different proteomics tools such as 2-D PAGE and mass<br />

spectrometry. The poteins can be identified either directly from the gels or through gels free<br />

approaches from the mixtures in solution (Figure 4) (79). The tandem mass spectrometry and<br />

MALDI-ToF-MS can be used from gel free and gel based protein identifications, respectively.<br />

The gel based identification approach has been reported many times and have several<br />

advantages over gel free approach (106, 109, 111). The gel electrophoresis can remove low<br />

16


Chapter 1<br />

molecular weight impurities and buffer ions during electrophoresis for which the MALDI-<br />

ToF has more sensitivity. Another advantage of the gel based approach is that polyacrylamide<br />

materials act like safe containers to handle and excise very small amount of the proteins. The<br />

protocols for the gel based approach are well established in comparison to gel free approach<br />

(80, 81). The mass spectrometers consist of three essential parts such as ionizer, analyzer and<br />

detector. The ionization source converts molecules into gas-phase ions. This ionization is<br />

facilitated by the energy provided by laser. As ions exit the ion source, the individual mass to<br />

charge (m / z) ratios are separated by a second device, a mass analyzer. The mass analyzer<br />

separates ions on the basis of mass to charge ratios. In MALDI-TOF-MS, the molecules are<br />

given mass values based on their movement or time of flight in the drift tube. Lighter ions fly<br />

faster than heavier ions and when it reaches to the third part of the mass spectrometer called<br />

detector. The signals are recorded on the detector and a mass spectrum is obtained. The<br />

detector records the charge produced during the hitting of ions on the surface. The spectrum is<br />

produced based on mass charge values. The experimental masses can be then compared to the<br />

theoretical masses available in the proteome of specific organism (80). The peptide mass<br />

fingerprinting approach can be used for searching protein databases. The principle behind<br />

PMF is very simple; the identified protein is the protein has the best match between the<br />

experimental and theoretical mass data. A theoretical digest of all the proteins sequences in<br />

the database by a specific enzyme can be used for predicting the individual peptides. These<br />

theoretical peptides can be searched to match peptides derived from the experimental mass<br />

data and then ideally one protein has to match the protein of interest. The protein will be<br />

considered as identified if the score is greater than the score fixed by Mascot with a p value<br />

less than 0.05. In addition, the experimental and theoretical values of the molecular weight<br />

and isoelectric point of a protein must be in the similar range. The databases such as NCBI<br />

and Swiss Prot can be searched for the identifications of proteins (82).<br />

17


Chapter 1<br />

Figure 4: The schematic presentation of protein identification using mass spectrometry<br />

(Adapted from Ref. 79).<br />

1.7. The in silico Approach<br />

The term in silico is actually the computational models or simulations based on the<br />

experimental data which can be used to make hypothesis and to predict about any biological<br />

process. The in silico approach has been widely used in pharmacology to make predictions,<br />

suggest hypothesis and ultimately leads towards new findings in drug designing and<br />

therapeutics. The in silico approach have the capability to increase the chances of discovery<br />

and decrease the laborious work in the laboratory. Many discoveries in the pharmacology<br />

were made by using the in silico approaches (83, 84, 85). It was stated that successful<br />

industrial companies are those that collect information from their experiments and introduce<br />

some rules that can show the pathways to obtain the target product for the clinical analysis<br />

18


Chapter 1<br />

followed by its introduction into the market. It has been reported that in silico approaches can<br />

help to make decisions and simulate about certain pharmaceutical processes which can further<br />

reduce the space between pharmaceutical and engineering based disciplines (84). To develop<br />

the computational models, it is necessary to have enough experimental data in the databases.<br />

These computational models can be exploited to predict about any biological process. The in<br />

silico approach has been reported to play its role as a complement to in vitro and in vivo<br />

experimentation (83). The upstream and downstream processes need to be improved utilizing<br />

the in silico approaches. The advances in the upstream processes such as fermentation have<br />

been already achieved and a product can be produced in several grams instead of milligrams<br />

due to the introduction of genetic engineering technologies (85). However, the major<br />

bottleneck still exists in the downstream processing of bioproducts. The current lack of the<br />

accurate data about the bioseparation of the products and the structural implications of the<br />

bioproducts are the main hurdles to design the purification models. A downstream approach<br />

has been proposed to collect the experimental data to design models for the improvement of<br />

the in silico approach (Figure 5). To collect the experimental data using modern mass<br />

spectrometry based proteomics tools <strong>after</strong> crude fractionation are highly required to design<br />

efficient bioprocesses (86). However, the in silico approach has been rarely reported in the<br />

downstream processing of proteins and the reason could be the lack of information with the<br />

real expression systems. Mostly model proteins were used to produce data and that data was<br />

then used to predict about the retention behavior of the other model proteins using<br />

mathematical models. The next step was then to compare the experimental retention time with<br />

that of predicted ones. In case of deviations from the predicted one, certain parameters have to<br />

be changed accordingly to calibrate the computer model. Once the model has been calibrated,<br />

no further investigations are required and the researcher can simulate the separation behavior<br />

of proteins on computer.<br />

19


Chapter 1<br />

Figure 5: Flow chart describing the fractionation and characterization procedure of the crude<br />

extract for obtaining the data about bioprocess parameters. The data produced can be stored<br />

in databases to design the purification models (Adapted from Ref. 86).<br />

Several researchers have reported different methodologies to predict about the retention time<br />

of model proteins (15, 87). However, the prediction models were based on the data collected<br />

from very few experiments and due to this reason there is still no universal scheme which can<br />

be considered as the best one for the downstream processing. These models have several<br />

drawbacks which can raise questions on their applicability. A major drawback was that the<br />

experimental data for these models has been derived from the model proteins instead of a host<br />

cell proteome. Another major drawback was that the experimental data was based on the<br />

retention time and / or volume of the model proteins. This is not applicable to a cell proteome<br />

because at a certain time during chromatography, several proteins are going to be eluted<br />

instead of a single protein. The shortcomings in the previous in silico approaches have<br />

motivated our group to report a paradigm shift from focussing on the purification of a target<br />

20


Chapter 1<br />

protein towards purification of the main protein contaminants in a host cell proteome. The use<br />

of the cell proteome in relation with the chromatographic method will produce more authentic<br />

and realistic data and can be stored in databases. These databases can be used to generate the<br />

purification models for easy in silico downstream processing of proteins (19). Several<br />

conclusions have been made in that work although the approach was not complete in its<br />

essence. The mass spectrometry and computational methods were missing in that report and<br />

the data reported was not enough to design a model. In this work, a complete picture has been<br />

presented ranging from chromatographic fractionation of the yeast cell proteome to the<br />

characterization of proteins at molecular level. The proteins properties have been used to<br />

correlate proteins at the molecular level to their separation behavior during HIC.<br />

1.8. Goal of the work<br />

Complementing all the above discoveries made by several authors, the current work further<br />

progressed to a more comprehensive understanding of the hydrophobic adsorbents and their<br />

effects on the separation behavior of a cell proteome during HIC. The adsorbents are usually<br />

composed of ligands and base supports. The comparison of the adsorbents (different ligands)<br />

hydrophobicity has been reported before with model proteins. However, the adsorbents<br />

hydrophobicity has been never studied comparatively with a proteome wide approach. The<br />

reason could be the experimentally demanding and laborious approach which no one ventured<br />

to adopt for their studies. Therefore, this study was planned for the comparative analysis of<br />

the hydrophobic adsorbents as a function of yeast cell proteome and in relation to protein<br />

properties.<br />

Under the frame of the current research work, the following objectives were targeted.<br />

• To explore the influence of different hydrophobic ligands in HIC as a function of<br />

complex cell proteome.<br />

21


Chapter 1<br />

• To uncover for the first time the influence of different base supports in HIC with a<br />

proteome wide approach.<br />

• To explore the physicochemical properties of proteins for their correlation with<br />

chromatographic separation and ability to differentiate among different ligands and<br />

base supports.<br />

• To reconnoitre the contaminant profile of the yeast cell proteome <strong>after</strong><br />

chromatographic fractionation using 2-D PAGE, mass spectrometry and<br />

bioinformatics.<br />

• To collect the experimental data and define certain operational windows, which can be<br />

annotated for the future implementation of the in silico downstream processing.<br />

22


Chapter 2<br />

EXPERIMENTAL APPROACH


Chapter 2<br />

2. Experimental approach<br />

2.1. Materials<br />

Yeast extract, peptone, dextrose, analytical grade buffer salts was purchased from Applichem<br />

(Darmstadt, Germany). The protease inhibitor cocktail was from Sigma-Aldrich Chemie<br />

GmbH (Steinheim, Germany). Acrylamide and dithiothreitol (DTT) were purchased from<br />

Carl-Roth GmbH (Karlsruhe, Germany). The Bradford Protein Assay Kit was purchased from<br />

Pierce Biotechnology Inc. (Rockford, IL, USA). ColorPlus pre-stained protein marker (New<br />

England Biolabs, Frankfurt, Germany) and N, N, N, N-tetramethyl-ethylenediamine (TEMED)<br />

were obtained from Serva (Heidelberg, Germany). The Toyopearl ®<br />

adsorbents were<br />

purchased from Tosoh Biosciences GmbH (Stuttgart, Germany). The C10/10<br />

chromatography column, chromatographic materials Sepharose-Phenyl, Source-Phenyl and<br />

pre-cast gels for electrophoresis were purchased from GE healthcare Europe GmbH (Munich,<br />

Germany).<br />

2.2. Experimental strategy<br />

The experimental strategy has been presented in Figure 6, which started with the cultivation<br />

of yeast cells followed by cells disruption. The cell proteome was collected <strong>after</strong> cell<br />

disruption and subjected to chromatographic fractionation using several hydrophobic<br />

adsorbents. The chromatographic fractions were collected in triplicate. The two dimensional<br />

gels were performed for the 42 fractions obtained during chromatography. Several proteins<br />

from each gel were selected for identification by MALDI-ToF-MS. Computational tools were<br />

used to calculate several properties of the proteins. Finally, the physicochemical properties of<br />

the proteins were correlated with the chromatographic behavior during HIC. All these<br />

experimental steps have been explained sequentially in the following text.<br />

24


Chapter 2<br />

Yeast cell proteome and chromatographic fractionation by HIC<br />

Phenyl ligated media (3) Polymethacrylate based media (3)<br />

21 fractions (7 each) 21 fractions (7 each)<br />

Perform 2-D PAGE for<br />

21 fractions separately<br />

Perform 2-D PAGE for 21<br />

fractions separately<br />

Identification of certain<br />

proteins from each gel<br />

by MALDI-ToF-MS.<br />

Identification of certain proteins<br />

from each gel by MALDI-ToF-<br />

MS.<br />

Bioinformatics approach<br />

Calculate several protein properties from the primary structure<br />

Average surface hydrophobicity calculated from three dimensional<br />

structure using MATLAB and STRIDE as computational tools<br />

Correlate protein properties with the chromatographic behavior and investigate<br />

the effects of ligands and base supports during chromatography<br />

Figure 6: Experimental strategy of the overall work.<br />

25


Chapter 2<br />

2.2.1. Cell Cultivation<br />

S. cerevisiae pBIVU02-1 (Biomedal, Sevilla, Spain) strain was streaked onto an YPD agar<br />

plate (1% Yeast extract, 2% Peptone, 2% Dextrose, 2% Agar) such that isolated single<br />

colonies will grow. The plates were incubated for two days. Then 30 ml of YPD liquid media<br />

was inoculated in a 250 ml flask with a single colony of cells from the YPD agar plate and<br />

grow overnight at 30°C with vigorous shaking (220 rpm) in the YPD liquid medium, pH 7.5.<br />

Then pre-culture was used to inoculate 300 ml YPD liquid media (1% yeast extract, 2%<br />

peptone, 1%w/v dextrose, pH 5.8) in 1.0 L baffled flasks. Cultivation was performed at 30°C,<br />

220 rpm, and for 48 hours. The cells were harvested by centrifugation in the late exponential<br />

growth phase at 3220 g for 15 minutes. The resulting cell pellet was washed three times with<br />

20 mM sodium phosphate buffer (pH 7.5) before cell disruption.<br />

2.2.2. Preparation of the crude extract<br />

Cell disruption was performed by bead milling at 4°C utilizing 0.5 mm glass beads. The cell<br />

paste (30 g) was suspended in 60 ml 20 mM sodium phosphate buffer (pH 7.5) and contacted<br />

by the grinding media for 30 min. One hundred micro litres of protease inhibitor cocktail (P<br />

8849; Sigma, St. Louis, MO, USA) was added per gram of cells. Cellular debris was<br />

separated by centrifugation at 3220 g for 25 min. The supernatant was filtered employing a<br />

0.45 µm membrane (Ministart Sartorius, Gottingen, Germany). The lysate was then ready to<br />

be used as a sample for chromatography experiments.<br />

2.2.3. Chromatographic fractionation<br />

Chromatographic fractionation was performed in an AKTA FPLC system equipped with a<br />

Frac-900 fraction collector and UNICORN 4.10 software for data collection and analysis (GE<br />

healthcare Europe GmbH, Munich, Germany). Two sets of chromatography stationary phase<br />

materials were used; one set had three different base supports (Sepharose-Phenyl, Source-<br />

Phenyl and Toyopearl-Phenyl) harbored the same ligand and the other set had different<br />

26


Chapter 2<br />

ligands (Toyopearl-Hexyl, Toyopearl-Butyl and Toyopearl-Ether) with same base support.<br />

Bed volume was 5.0 ml; the quality of the packing was evaluated by residence time<br />

distribution analysis using 1% acetone as a tracer (88). Mobile phases were prepared as<br />

follows: buffer A: 20 mM sodium phosphate buffer pH 7.5 containing 1.6 M (NH 4 ) 2 SO 4 , and<br />

buffer B: 20 mM sodium phosphate buffer. Buffers were filtered and degassed prior to use.<br />

After equilibration in buffer A (10 CV), the cell lysate (10 ml; 8.0 mg/ml) was applied to the<br />

column. The sample was introduced by a 10 ml super loop (GE health care, Uppsala, Sweden).<br />

Non-retained materials were washed out in buffer A (5 CV) and applying a linear gradient (25<br />

CV) from 0% to 100% of buffer B exerted elution. The flow rate was 1 mL / min. Fractions<br />

were collected utilizing the following salt concentration values as cut-off: 1 (1.6 M), 2 (1.4<br />

M), 3 (1.0 M), 4 (0.8 M), 5 (0.6 M), 6 (0.2 M) and 7 (0 M). The chromatographic experiments<br />

were performed in triplicate and the fractions were pooled before analysis.<br />

2.2.4. Membrane dialysis<br />

The materials collected in each of the seven fractions were further desalted by membrane<br />

dialysis (Pierce Rockford, IL, United States). The membranes (7 KDa) with samples were<br />

stirred in 5 litres bottle full of water. The salt buffer ions have been removed from the samples<br />

by diffusion. The dialysis was performed for two days in a cold room (4°C) with a renewal of<br />

fresh water <strong>after</strong> each 12 hours interval. After desalting, the samples were concentrated under<br />

vacuum (Vacufuge concentrator 5301, Eppendorf AG, Hamburg, Germany).<br />

2.2.5. Two dimensional polyacrylamide gel electrophoresis (2-D PAGE)<br />

The chromatography fractions were obtained as described earlier and dissolved in Lysis<br />

buffer (7 M urea, 4% w/v CHAPS, 2 M thiourea, 2% v/v Pharmalyte TM 3-10, and 1% w/v<br />

DTT). The samples were then centrifuged (3220 g, 15°C) to remove any insoluble material.<br />

Isoelectric focussing was performed using 7 cm Immobiline TM Dry strips pH 3-10, kept at<br />

20°C in a flatbed Multiphor II unit (GE Healthcare Europe, GmbH, Munich, Germany). The<br />

27


Chapter 2<br />

IPG strips were kept for 16 hours in rehydration buffer (6 M urea, 2 M thiourea, 1% CHAPS,<br />

0.4% DTT, and 0.5% v/v pharmalyte TM 3-10). The sample 200 µl (65 µg proteins in 75 + 125<br />

µl of rehydration buffer) was applied in in-gel rehydration buffer instead of cup loading. The<br />

sample was allowed to absorb on a strip for 15 min, and then covered with silicon oil. After a<br />

sixteen hour rehydration period, focussing was performed. IEF was performed on these<br />

conditions: 50 V, 120 min. 150 V, 90 min. 500 V, 60 min, 1500 V, 40 min. 2500 V, 40 min<br />

and 3500 V, 180 min at 20°C. The minute amount of buffer ions present in the samples was<br />

removed by applying low voltage (100 volts) at the beginning for 5 hours and the filter paper<br />

beneath the electrode (where the salt ions have collected) was changed time by time. After<br />

focussing, the strips were first equilibrated for 15 minutes in 10 ml equilibration buffer (50<br />

mM Tris-HCl, pH 8.8, 6 M urea, 30% glycerol, 2% w/v SDS and 10 mg/ml DTT) containing<br />

20 µl bromophenol blue and subsequently for additional 15 minutes in the same buffer<br />

containing 25 mg/ml iodoacetamide instead of DTT. After equilibration, the strips were kept<br />

on top of a 12.5% SDS vertical slab and covered by 0.7% hot low melting point agarose in<br />

SDS electrophoresis running buffer (25 mM Tris, 192 mM glycine, 0.1% SDS). The sodium<br />

dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) was performed in a Hoefer<br />

mini VE vertical electrophoresis system in 1 mm thick 10 cm × 10.5 cm gels, 25 mA per gel<br />

and 120 volts (19). Colloidal coomassie blue was used for staining the gels (89). The gels<br />

were first fixed for at least three hours in the fixing solution (50% ethanol and 3% phosphoric<br />

acid). The gels tend to shrink during fixing. Then the gels were washed in deionized water<br />

three times for 20 minutes at a time. After washing, the gels were preincubated for 1 hour in<br />

staining solution (34% methanol, 3% phosphoric acid and 17% w/v ammonium sulfate). Then<br />

a small amount of coomassie blue was added with a spatula tip to the gels and stained<br />

overnight. After staining, the gels were washed with deionized water to remove the<br />

background stain. The gels were scanned by 48 bit Epson color scanner (Epson perfection<br />

4990 Photo). 2-D gels of the chromatography fractions were performed in duplicate and the<br />

28


Chapter 2<br />

number of spots was counted based on size and isoelectric point values. Certain spots were<br />

selected for in-gel digestion and protein identification.<br />

2.2.6. Protein content of the fractions<br />

The protein concentration in the fractions was determined by the Bradford protein assay kit<br />

(Pierce, Rockford, IL) as per the manufacturer's instructions. The protein content of the crude<br />

extract was measured before using it for chromatographic experiments. The protein content<br />

was also measured in all the chromatographic fractions collected during HIC.<br />

2.2.7. Protein identification by MALDI-ToF-MS<br />

Certain protein spots were manually excised from the 2-D gel for the identification purposes.<br />

In-gel digestion was performed according to the Shevchenko´s protocol (90) with some<br />

modifications. A high amount of salt was present in the samples, so an additional step was<br />

followed where a wash solution (50% v/v methanol and 5% v/v acetic acid) was added to gel<br />

pieces and kept overnight. The washing step was repeated three times. The wash solution was<br />

very effective in removing all kinds of impurities hindering the identification process. The<br />

other additional step was the use of 0.5% TFA to dissolve the dried peptides instead of 0.1%<br />

TFA mentioned in Shevchenko´s protocol. The reason was the presence of minute amount of<br />

buffer ions in the samples which was negatively affecting the spotting of proteins on the<br />

MALDI target plate. After washing with 0.5% TFA, the spots on the plate were solid and<br />

resulted in mass spectra of high quality. The modifications in the existing protocol should be<br />

considered in the protein identifications specific in case when HIC is involved. The overall ingel<br />

digestion process was performed in seven steps such as washing, de-staining, reduction,<br />

alkylation, trypsin digestion, extraction and concentration of the peptide extract (Table 2).<br />

29


Chapter 2<br />

Table 2: The in-gel digestion protocol optimized specifically for the proteins fractionated by<br />

HIC<br />

Steps<br />

Description<br />

1. Washing the gels The gel pieces were washed overnight, with 50% (v/v) methanol<br />

and 5% (v/v) acetic acids for the removal of salts and other<br />

impurities.<br />

2. De-staining After washing, the gel pieces were incubated in 200 µl of 100<br />

mM ammonium carbonate/acetonitrile (1:1 v/v) for 30 minutes<br />

to destain the gels.<br />

3. Reduction After destaining, 50 µl of 10 mM DDT dissolved in 100 mM<br />

ammonium carbonate were added to gel pieces and incubated<br />

for 45 minutes at 56°C to reduce the disulfide bridges.<br />

4. Alkylation Then 100 µl of 55 mM iodoacetamide was added to samples and<br />

incubated for 20 minutes in the dark to inhibit the reformation of<br />

disulfide bonds and alkylate the free cysteins.<br />

5. Trypsin digestion Then 10 µl of trypsin and 90 µl of 25 mM ammonium<br />

bicarbonate in 9% acetonitrile solution were added to each<br />

sample and incubated in ice for 3 hours to absorb maximum<br />

enzymes. The samples were then incubated for 16 hours at 37°C<br />

in the digestion buffer for protein digestion.<br />

6. Extraction of<br />

peptides<br />

7. Concentration of<br />

the peptide extract<br />

a. 200 µl of Milli Q water was added to each sample and<br />

incubated for 15 minutes at 37°C to extract hydrophilic<br />

peptides.<br />

b. 200 µl of extraction buffer (1:2 v/v 5% formic<br />

acid/acetonitrile) was added to each sample and<br />

incubated for 15 minutes at 37°C to extract hydrophobic<br />

peptides.<br />

The peptide extract was concentrated and 15 µl of 0.5% v/v<br />

TFA was added to each sample, followed by centrifugation at<br />

10,000 rpm for 10 minutes. The samples were then ready for<br />

identification by MALDI-ToF-MS.<br />

After digestion, the sample was collected as supernatant and 3 µl of the sample was spotted<br />

on the MALDI target plate followed by 0.5% TFA for washing. The sample was air dried at<br />

room temperature. All mass spectra were calibrated using peptides from the auto digestion of<br />

trypsin. The peak lists were searched using the Mascot search engine<br />

30


Chapter 2<br />

(http://www.matrixscience.com) against NCBI and SwissProt databases. The search<br />

parameters were carbamidomethylation and methionine oxidation entered as variable<br />

modifications and 1 missed trypsin cleavage site. In all protein identifications, the scores were<br />

greater than the score fixed by Mascot as significant with a p value < 0.05 (91). The identified<br />

proteins were further characterized and analyzed with computational tools.<br />

2.2.8. Characterization of proteins utilizing Bioinformics<br />

The same procedure has been followed for the calculation of ASH as described by J.C.<br />

Salgado et al utilizing the Cowan-Whittaker scale (32, 92). Average surface hydrophobicity<br />

(ASH) of the proteins was calculated by writing a package in MATLAB, where it was<br />

mandatory to use the three-dimensional structure of a protein collected from protein databank<br />

(PDB, http://www.rcsb.org/pdb). STRIDE was used to generate a text file, which can<br />

determine the accessible surface area (ASA) of each amino acid at a particular position. After<br />

calculating the ASA, the ASH was measured by the package in MATLAB, programmed with<br />

the normalized values of Cowan-Whittaker hydrophobicity scale. According to this method,<br />

the ASH can be calculated by equation 1 as follows;<br />

Φ = ˆ rφ<br />

(1)<br />

suface<br />

∑<br />

i<br />

i∈A<br />

i<br />

Where<br />

Φ<br />

surface is the ASH of a protein, A is the collection of the 20 possible amino acids<br />

and ( φ i ) is the hydrophobicity of the amino acids of type i. The hydrophobicity of each amino<br />

acid was assigned according to Cowan-Whittaker scale. The fraction of superficial area ( r ∧ i)<br />

occupied by the amino acid i can be calculated by equation 2 as follows;<br />

31


Chapter 2<br />

∧<br />

r<br />

i<br />

=<br />

Si<br />

S<br />

∑<br />

jεA<br />

j<br />

(2)<br />

In equation (2),<br />

Si<br />

is the sum of the accessible surface area (ASA) for all the amino acids of<br />

type i. S j is the total surface area of the protein. The r ∧ i is the fraction of surface area occupied<br />

by amino acid i and A is the collection of 20 possible amino acids (93). The concept of this<br />

approach was based on the assumption that each amino acid on the protein surface contributes<br />

regarding its abundance, with the properties associated to the protein surface (32, 67).<br />

Average hydrophobicity (AH), average polarity, average bulkiness and average flexibility<br />

(AF) were calculated from the primary structure by using the scales available at Expasy<br />

ProtScale tool (94). The average hydrophobicity values were measured employing three<br />

different scales i.e. Tanford (T), Cowan-Whittaker (CW) and Miyazawa-Jernigan (MJ),<br />

respectively (25, 92, 95). The average bulkiness was calculated by the scale proposed by<br />

Zimmerman (96). The polarity values were calculated by adopting the Grantham’s and<br />

Zimmerman’s scales (96, 97). The flexibility indices introduced by Bhaskaran and<br />

Ponnuswamy were used to calculate the flexibility values of proteins (98).<br />

32


Chapter 3<br />

RESULTS AND DISCUSSION


Chapter 3<br />

3. Results and Discussion<br />

3.1. Influence of the different ligand chemistries during hydrophobic<br />

interaction chromatography<br />

3.1.1. Summary<br />

Several reports are available in literature in which various ligands have been compared based<br />

on the retention time of model proteins in hydrophobic interaction chromatography (HIC) (62,<br />

87, 99). However, the different ligands were comparatively investigated for the first time with<br />

the proteome wide approach. A proteome wide approach has direct industrial applicability and<br />

has several advantages over work performed with model proteins. This work is in the<br />

continuation of a previous report from our group, where the ion exchange beads were used to<br />

investigate the separation behavior of the insect cell proteome (19). Several conclusions were<br />

made in that report, although the approach was not complete in its essence. The mentioned<br />

approach was extended in this work for the comparative studies of the hydrophobic ligands<br />

employing some additional steps of mass spectrometry and bioinformatics; which were<br />

missing in the previous approach. This approach analyzed the chromatographic separation of<br />

the overall protein contaminants in the host cell proteome instead of focussing on any of the<br />

target protein. The Saccharomyces cerevisiae cell proteome was fractionated by different<br />

ligands during HIC. The chromatographic fractions were collected at specific salt<br />

concentrations, followed by their separation on two-dimensional polyacrylamide gel<br />

electrophoresis (2-D PAGE). Several spots were selected from each 2-D gel and subjected to<br />

identification by Matrix Assisted Laser Desorption/Ionization Time-of-Flight Mass<br />

Spectrometry (MALDI-ToF-MS). Later on, several protein properties such as average surface<br />

hydrophobicity, average hydrophobicity, average flexibility, average bulkiness and average<br />

polarity were calculated for the identified proteins using different bioinformatics tools. These<br />

properties were investigated for the first time with the proteome wide approach and correlated<br />

35


Chapter 3<br />

to the retention behaviors in HIC. These properties were also studied for their ability to<br />

differentiate among the three different ligands. In addition, this work was planned to collect<br />

the experimental data from the three ligands which can be used to design the purification<br />

model of recombinant proteins utilizing the in silico approach.<br />

3.1.2. Scientific background of the approach<br />

3.1.2.1. Hydrophobic adsorbents<br />

Hydrophobic interaction chromatography (HIC) is an important chromatographic technique<br />

used for the industrial scale purification of proteins. HIC was preferred over other techniques<br />

such as affinity chromatography, ion exchange chromatography and reverse phase<br />

chromatography, due to the mild hydrophobic interactions. These weaker interactions<br />

minimize the unfolding of the proteins and maintain the biological activities (68, 100). The<br />

hydrophobic adsorbents and protein properties are the two main parameters affecting the<br />

chromatographic behavior during HIC. The hydrophobic adsorbents are usually composed of<br />

hydrophobic ligands attached to the functional groups of the base supports. The hydrophobic<br />

ligands are usually ether, butyl, hexyl and phenyl groups depending on their chain lengths. As<br />

the chain lengths of the ligands are increasing, the water exclusion potential of the ligands is<br />

also increasing. In other words, the chain length of the ligand has a direct relationship to<br />

hydrophobic interactions. If the hydrophobic interactions between proteins and adsorbents are<br />

higher, the proteins will be tightly bound to the column and will elute at the end during<br />

chromatography. The hydrophobicity studies of the different ligands have been provided by<br />

the Vendor company and also reported in the open literature (62, 87, 99, 101). The ether<br />

ligand was reported to be the most hydrophilic and hexyl as the most hydrophobic ligand as<br />

presented in Figure 7 (99).<br />

Toyopearl-Ether < Toyopearl-Butyl < Toyopearl-Hexyl<br />

36


Chapter 3<br />

Ligand<br />

Base support<br />

(Toyopearl®)<br />

Figure 7: The chemistry and comparison of different ligands for their hydrophobicity based on<br />

model protein retention data (Adapted from Ref. 99).<br />

The ligands are usually attached to toyopearl ®<br />

base support by glycidyl ether coupling<br />

procedures. This procedure introduces a short spacer arm to minimize steric hindrance. At the<br />

start, HIC was a slow technique; rapid separation has been made possible by the introduction<br />

of micro particulate semi rigid supports (Toyopearl®) (101). This revealed the importance of<br />

toyopearl adsorbents at the industrial level. The physicochemical properties of the adsorbents<br />

are given in Table 3 (99). The three hydrophobic adsorbents of 65-100 µm bead sizes were<br />

selected for this study, which is the most suitable range for industrial scale purification. The<br />

hydrophobic charm of the ligands was narrated before with toyopearl as a common base<br />

support. However, the previous reports were based on the retention time of few model<br />

proteins and have been never studied with the biological mixtures. Hence, it was planned to<br />

37


Chapter 3<br />

investigate the hydrophobicity of the following three ligands as a function of the cell<br />

proteome and in relation to protein properties (Figure 8).<br />

Table 3: Physical properties of the adsorbents with different ligands (99)<br />

Adsorbents<br />

Base support<br />

Ligands<br />

Beads size<br />

(µm)<br />

Pore size<br />

(nm)<br />

DBC a (mg/mL)<br />

Toyopearl-<br />

Butyl<br />

polymethacrylate Butyl 100 100<br />

32<br />

Toyopearl-<br />

Hexyl<br />

polymethacrylate Hexyl 100 100<br />

33<br />

Toyopearl-<br />

Ether<br />

DBC a<br />

polymethacrylate Ether 65 100<br />

Dynamic binding capacity<br />

12<br />

Yeast proteome separation<br />

on hydrophobic adsorbents<br />

(HIC)<br />

Comparison of<br />

three different<br />

ligands<br />

Comparison of<br />

three base<br />

supports<br />

Protein<br />

characteristics<br />

Toyopearl-Hexyl<br />

Toyopearl-Butyl<br />

Toyopearl-Ether<br />

Source-Phenyl<br />

Sepharose-Phenyl<br />

Toyopearl-Phenyl<br />

Protein<br />

hydrophobicity and<br />

others<br />

Figure 8: The comparison of the three different ligands with common base support based on<br />

protein properties and as a function of the expression system.<br />

38


Chapter 3<br />

3.1.2.2. Protein properties in the context of HIC<br />

The physicochemical properties of the proteins such as average hydrophobicity, average<br />

polarity, average flexibility and average bulkiness can be determined by different scales<br />

proposed by several researchers (Table 4). These average properties were based on the<br />

exposed and buried amino acid residues of a protein. The average hydrophobicity is the<br />

average of the hydrophobicities of the exposed and buried amino acids of a protein. The<br />

average hydrophobicity of the proteins can be measured by the scales proposed by different<br />

researchers (25, 92, 95). Several experimental and theoretical approaches were used to<br />

determine the hydrophobicity indices for the individual amino acids. These methods have<br />

assigned different hydrophobicity values for each amino acid and can be used as different<br />

amino acid hydrophobicity scales. The scale of Cowan-Whittaker was derived from the amino<br />

acid retention times during chromatographic experiments by RP-HPLC (92). The scale of<br />

Miyazawa-Jernigan was based on the contact energies estimated from the protein tertiary<br />

structures. A direct relationship was observed between the contact energies of adjacent<br />

residues in the crystal structure and their corresponding hydrophobicities reported by Tanford<br />

(25, 95). The theoretical contact energies between amino acid residues in the crystal structures<br />

were used to assign the hydrophobicity value to each amino acid. The amino acid<br />

hydrophobicity scale proposed by Tanford was based on the amino acids solubility in water as<br />

well as in spontaneously increasing concentration of organic solvent (ethanol). The amino<br />

acids were classified as hydrophobic, intermediate and hydrophilic based on the free energy<br />

change by the transfer of amino acids from ethanol to water. Based on the data, a<br />

hydrophobicity scale was proposed for amino acid residues, where their free energy transfer<br />

values become more and more positive as the hydrophobic character of the protein increases<br />

(102). This revealed that hydrophobic amino acids will tend towards organic phase while<br />

hydrophilic amino acids will prefer to stay in the water. In other words, charged residues of<br />

the protein will tend towards the surface and in contrast, the hydrophobic residues will stay in<br />

39


Chapter 3<br />

the interior of the protein (25). The average hydrophobicity parameter was reported in<br />

literature for its relation to protein structure (66). However, this parameter was rarely<br />

investigated in context of the HIC. Average flexibility is another property which can be<br />

calculated by using Bhaskaran and Ponnuswamy’s scale. Average flexibility represents the<br />

average fluctuations of the all residues in a given protein. It was reported that proteins with<br />

high flexibility will have more chances to unfold and expose hydrophobic residues on the<br />

protein surface. This will increase hydrophobic interactions between proteins and adsorbents<br />

and result in longer retention of proteins during chromatographic experiments (33).<br />

Table 4: The physicochemical properties of the proteins investigated in this study<br />

Protein properties Scales Bioinformatic tools Source<br />

Average<br />

hydrophobicity<br />

Cowan-Whittaker’s<br />

scale,<br />

Expasy ProtScale tool Primary<br />

structure<br />

Miyazawa-Jernigan’s<br />

scale<br />

Tanford’s scale<br />

Average flexibility Bhaskaran and<br />

Ponnuswamy’s scale<br />

Expasy ProtScale tool Primary<br />

structure<br />

Average polarity Grantham’s scale,<br />

Zimmerman’s scale.<br />

Expasy ProtScale tool Primary<br />

structure<br />

Average bulkiness Zimmerman’s scale Expasy ProtScale tool Primary<br />

structure<br />

Average surface<br />

hydrophobicity<br />

Cowan-Whittaker’s<br />

scale<br />

Expasy ProtScale tool,<br />

MATLAB and STRIDE<br />

softwares<br />

Three<br />

dimensional<br />

structure<br />

The polarity scales proposed by Grantham and Zimmerman were used to quantify the polarity<br />

based on the protein sequence (96, 97). All the above mentioned properties were based on the<br />

primary structure of a protein. However, ASH was the most remarkable property in the<br />

40


Chapter 3<br />

context of HIC based on the exposed amino acid residues of a three dimensional protein<br />

structure. The ASH of a protein has a direct proportionality to the exposed hydrophobic<br />

surface area on the protein surface. The three crystal structures identified in this work have<br />

been compared for their hydrophobic surface residues and elution position during<br />

chromatography (Figure 9). The crystal structures of the enzymes were obtained using<br />

chimera as a visualization tool (http://www.cgl.ucsf.edu/chimera). The protein such as Serine<br />

threonine protein kinase (A) has very less hydrophobic surface residues in comparison to the<br />

protein Ubiquitin carboxyl terminal hydrolase 4 (B). Due to this reason, the protein A was<br />

eluted earlier than protein B during chromatography. In the same way, the protein B has less<br />

hydrophobic surface residues as compared to the protein Phosphogycerate mutase 1 (C). Due<br />

to the less hydrophobic residues, the protein B was eluted earlier than protein C. There are<br />

several other important enzymes identified at different elution ranges and can be analyzed at<br />

the molecular level for their hydrophobic content and corresponding chromatographic<br />

behavior. These surface hydrophobic residues can be quantified in terms of ASH from the<br />

three dimensional structure of a protein utilizing MATLAB and STRIDE as computational<br />

tools. ASH has been widely exploited to predict the retention times of model proteins during<br />

HIC (103). However, all the above mentioned properties have been never studied in relation<br />

to the host derived proteins. Therefore, the above mentioned properties were investigated for<br />

their correlation with the chromatographic behavior and potential to differentiate among the<br />

ligands. This work was planned to provide the data from a real bioprocess for designing the<br />

purification model of the recombinant proteins utilizing an in silico approach.<br />

41


Chapter 3<br />

A<br />

B<br />

C<br />

Figure 9: The crystal structures of three proteins identified in this work. These proteins were<br />

eluted at different positions during chromatography depending on the surface hydrophobic<br />

residues. The red and blue colors represent the hydrophobic and hydrophilic surface residues,<br />

respectively.<br />

42


Chapter 3<br />

3.1.2.3. Towards an in silico approach for the downstream processing of bioproducts<br />

The in silico approach is actually the computational models or simulations based on the<br />

experimental data which can be used for making hypothesis and predictions about any<br />

biological process (83, 84). This approach has been widely used in the pharmaceutical<br />

industry for drugs discovery as discussed in chapter 1. However, the in silico approach was<br />

rarely evaluated in practical terms for the downstream processing of proteins. Several research<br />

groups had proposed models for the prediction of the chromatographic behavior of proteins<br />

onto hydrophobic adsorbents with varying degrees of success but none of these approaches<br />

have gained a universal acceptance (26, 87, 103). The first and major criticism was that all the<br />

schemes were based on model proteins instead of real biological crude extract. The second<br />

criticism was that how can one predict about the separation of any specific protein from<br />

biological mixtures based on the data of few homologous model proteins and without<br />

considering the chromatographic fractionation of the host cell proteome. The third criticism<br />

was that those models were based on the retention time and / or volume of the model proteins<br />

and these two parameters have no applicability in terms of real bioprocess. At any specific<br />

space of time (minute), several proteins are supposed to be eluted instead of a single protein<br />

during chromatographic fractionation of a cell proteome. Instead of retention time and / or<br />

volume, the salt concentration should be specified for the protein elution under specific<br />

chromatographic conditions. They never tried a model based on the chromatographic<br />

separation of the total cell proteome as the approach is experimentally demanding and<br />

laborious. This approach has been adopted in this work as presented in Figure 10, where<br />

several techniques were employed ranging from chromatographic fractionation of cell<br />

proteome to identification and analysis of the proteins available in each chromatographic<br />

fraction. The identified proteins were characterized by several protein properties and these<br />

properties were then further correlated to the chromatographic behavior. The protein<br />

properties were also investigated for their ability to differentiate among the adsorbents. Aside<br />

43


Chapter 3<br />

of the comparative analysis of the ligands as a main ambition, this work was planned to<br />

collect the experimental data based on the chromatographic separation of a cell proteome. The<br />

information obtained regarding proteins elution ranges on different adsorbents and the<br />

corresponding protein properties can be used for designing the purification models of the<br />

recombinant proteins by an in silico approach. This work has given a first draft of the<br />

experimental data to design models and predict the elution position of the recombinant<br />

proteins during HIC through computer simulations.<br />

Figure 10: Schematic representation of the methodology towards an in silico approach.<br />

44


Chapter 3<br />

3.1.3. Results and Discussion<br />

3.1.3.1. Chromatographic profiles of different ligands<br />

The S. cerevisiae PBIVU02-1 was chosen as an expression system, because it was the first<br />

organism to have its genome completely sequenced. Scouting chromatographic experiments<br />

were performed in order to evaluate the separation behavior of the S. cerevisiae cell proteome<br />

on different ligands (Ether, Butyl and Hexyl) employing Toyopearl® as a common base<br />

support. A decreasing gradient of ammonium sulfate was performed from 1.6 to 0 M and<br />

seven fractions were collected at specific salt concentrations. The salt concentrations of the<br />

fractions have been specified by numbers such as 1 (1.6 M), 2 (1.4 M), 3 (1 M), 4 (0.8 M), 5<br />

(0.6 M), 6 (0.2 M) and 7 (0 M). These salt concentrations at elution positions were defined for<br />

the first time in order to investigate the proteins available in each chromatographic fraction.<br />

The prediction models reported in the previous studies were based on the retention time and /<br />

or volume of the model proteins instead of the salt concentrations at elution stage (62).<br />

However, the retention time and / or volume have no applicability for the fractionation of a<br />

cell proteome. Several proteins are supposed to be eluted at a fraction of time (minute), so to<br />

assign a specific elution time to any protein from a biological mixture is impractical.<br />

Therefore, the fractions were collected on specific salt concentrations during chromatography.<br />

The chromatograms revealed no clear matching in the separation profiles of the three<br />

adsorbents for the same crude extract (Figure 11). A high resolution was observed for<br />

Toyopearl-Ether- than -Butyl and -Hexyl. The reason could be the smaller bead size of<br />

Toyopearl-Ether. Another reason could be that resolution decreases when the chain length is<br />

increasing i.e. Toyopearl-Hexyl- and -Butyl have higher chain lengths than -Ether (51). These<br />

results confirmed that bead size and chain length of the ligands have a significant role in<br />

resolution. Overall resolution of chromatographic separation was very low and the reason can<br />

be the high amount of sample which was applied in this micro preparative chromatography.<br />

45


Chapter 3<br />

Figure 11: Represent the chromatographic profiles obtained <strong>after</strong> the fractionation of the cell proteome by HIC using Toyopearl-Hexyl, Toyopearl-<br />

Butyl and Toyopearl-Ether. Proteins were loaded in a mobile phase composed of 1.6 M (NH 4 ) 2 SO 4 at pH 7.5. The elution was exerted for 25 CV. The<br />

fractions were collected in 7 groups utilizing the following salt concentrations as cut off values: 1 (1.6 M), 2 (1.4 M), 3 (1.0 M), 4 (0.8 M), 5 (0.6 M), 6 (0.2<br />

M) and 7 (0.0 M). Each fraction was collected by pooling in triplicate. (- Refers to 280 nm mAU, --- is Buffer B %).<br />

46


Chapter 3<br />

The applied protein load (40 mg) for chromatographic experiments was within the range of<br />

dynamic binding capacities of the adsorbents. A high amount of proteins was adsorbed on<br />

Toyopearl-Hexyl- and -Butyl which confirmed the high dynamic binding capacities of these<br />

adsorbents. Toyopearl-Ether has shown less binding affinity and most of the proteins were<br />

lost in flow through as non retained components. The results revealed the binding affinities of<br />

the adsorbents for a cell proteome. In case of Toyopearl-Ether, almost 80% of the total load<br />

was lost during washing as evidenced by the Bradford method. However, for Toyopearl-Butyl<br />

and -Hexyl almost 50% of the proteins were lost in the flow through. The remaining half of<br />

the proteins were retained in the column and eluted according to their hydrophobicity during<br />

chromatography. A high amount of proteins was lost during flow through which confirmed<br />

the less hydrophobic nature of the S. cerevisiae cell proteome (104). The results suggested the<br />

use of high salt concentrations in order to increase the adsorption affinity of the yeast cell<br />

proteome during HIC.<br />

3.1.3.2. Two dimensional gel electrophoresis and proteins identification<br />

The two dimensional gels were performed for all the chromatography fractions collected<br />

within a specific operational window. Several proteins were selected from each gel for<br />

identification purposes; however in a few cases the tiny and weakly stained proteins were<br />

identified. Four out of the total identified proteins were visible on the gels with naked eyes,<br />

however were not visible when photographed and it is sometimes the case with weakly<br />

stained proteins (Figures 12-14). There are several reports where weakly stained proteins<br />

were identified, but the spots were not visible on the 2-D gel pictures (105-111). In total,<br />

ninety six proteins were identified using MALDI-ToF-MS, which revealed the applicability of<br />

the in-gel digestion method optimized in this work specifically for the proteins fractionated by<br />

HIC (Tables 5-7). The identified proteins can further provide the contaminant profile of the<br />

yeast cell proteome and is an addition to the proteomics technology. In the case of Toyopearl-<br />

47


Chapter 3<br />

Ether, some 2-D gels were observed to have the restricted number of protein spots which<br />

could be used as an advantage to have the targeted product on the specific conditions. Overall,<br />

three dimensional separation of the yeast cell proteome was performed, based on<br />

hydrophobicity (HIC), and then followed by isoelectric point and molecular weight (2-D<br />

PAGE). The three dimensional characterization can be a suitable method when choosing an<br />

efficient host for the protein purification (112). The chromatographic methods can be used in<br />

the prefractionation steps for proteomics studies. Several chromatographies have been used<br />

previously to decrease the dynamic range of the cell proteomes on 2-D gels such as heparin<br />

and hydroxyapatite used for the Haemophilus influenza and Escherichia coli, respectively (76,<br />

109). However, it was an initial effort to fractionate the yeast cell proteome by several<br />

hydrophobic adsorbents followed by its proteomics analysis.<br />

48


Chapter 3<br />

Figure 12: Represent 2-D D gels performed for the samples fractionated by Toyopearl-Ether.<br />

Each gel represents the specific fractions collected at different salt ranges such as A (1.6 M), B<br />

(1.4 M), C (1 M), D (0.8 M), E (0.6 M), F (0.2 M) and G (0.0 M). The spots indicated by numbers<br />

were selected for identification. The encircled spot has been explained in section 3.1.3.2.<br />

49


Chapter 3<br />

Figure 13: Represent 2-D D gels performed for the samples fractionated by Toyopearl-Butyl.<br />

Each gel represents the fractions collected at different salt ranges such as A (1.6 M), B (1.4 M),<br />

C (1 M), D (0.8 M), E (0.6 M), F (0.2 M) and G (0.0 M). The spots indicated by numbers were<br />

selected for identification. The encircled spot has been explained in section 3.1.3.2.<br />

50


Chapter 3<br />

Figure 14: Represent 2-D D gels performed for the samples fractionated by Toyopearl-Hexyl.<br />

Each gel represents the fractions collected at different salt ranges such as A (1.6 M), B (1.4 M),<br />

C (1 M), D (0.8 M), E (0.6 M), F (0.2 M) and G (0.0 M). The spots indicated by numbers were<br />

selected for identification. The encircled spots have been explained in section 3.1.3.2.<br />

51


Chapter 3<br />

Table 5. List of the identified proteins fractionated by Toyopearl-Ether and their characterization with different properties<br />

Prot A Salt B Prot ≠ C Prot. I.D D Mr E pI F PDB G ASH H AH I AH J AH K AP L AP M AF N AB O<br />

Low 1.6 A1-TE RAD16_YEAST 91 8.54 3AUX 0.4225 0.403 0.503 0.587 0.475 0.315 0.603 0.656<br />

A2-TE OYE2_YEAST 44 6.13 3RND 0.4234 0.390 0.524 0.606 0.465 0.274 0.612 0.638<br />

A3-TE Fox2p 96 8.96 3QRF 0.4289 0.363 0.484 0.582 0.507 0.318 0.588 0.637<br />

A4-TE YN92_YEAST 69 6.15 3OQ0 0.4319 0.365 0.491 0.585 0.503 0.310 0.591 0.671<br />

1.4 B1-TE PRX1_YEAST 29 8.97 2PWJ 0.4370 0.402 0.492 0.587 0.501 0.271 0.598 0.675<br />

B2-TE SYF1_YEAST 100 5.17 3GIX 0.4396 0.372 0.528 0.608 0.500 0.290 0.593 0.667<br />

B3-TE TPIS_YEAST 26 5.74 3TH6 0.4411 0.408 0.503 0.597 0.454 0.302 0.599 0.688<br />

B4-TE APM2_YEAST 69 6.41 1HES 0.4436 0.402 0.507 0.596 0.476 0.316 0.603 0.667<br />

1.0 C1-TE UPP1 24 5.58 2EHJ 0.4451 0.382 0.511 0.605 0.478 0.279 0.597 0.662<br />

C2-TE Fur1p 23 5.84 2JBH 0.4458 0.385 0.512 0.620 0.476 0.280 0.598 0.667<br />

C3-TE CAT8_YEAST 160 9.13 3SNP 0.4505 0.389 0.514 0.612 0.472 0.289 0.608 0.636<br />

C4-TE BUR1_YEAST 74 9.55 3R22 0.4495 0.390 0.511 0.614 0.478 0.284 0.624 0.649<br />

Medium 0.8 D1-TE TRM11_YEAST 50 7.64 1ZJR 0.4516 0.411 0.517 0.616 0.462 0.288 0.612 0.639<br />

D2-TE FSF1_YEAST 35 9.84 3OJ2 0.4584 0.412 0.518 0.617 0.461 0.279 0.614 0.648<br />

D3-TE YO316_YEAST 7 11.66 2Y6V 0.4612 0.414 0.520 0.619 0.443 0.277 0.643 0.627<br />

D4-TE IMH1_YEAST 105 5.52 3TCX 0.4618 0.415 0.526 0.620 0.457 0.275 0.645 0.634<br />

D5-TE FAF1_YEAST 38 9.33 3R3M 0.4620 0.406 0.528 0.621 0.455 0.274 0.614 0.698<br />

0.6 E1-TE ICP55_YEAST 57 6.11 1VZ2 0.4644 0.414 0.527 0.607 0.456 0.273 0.615 0.641<br />

E2-TE CORO_YEAST 72 5.64 2AQ5 0.4660 0.417 0.531 0.622 0.448 0.264 0.636 0.629<br />

E3-TE SPC2_YEAST 20 9.22 3SBT 0.4684 0.426 0.539 0.619 0.435 0.285 0.628 0.684<br />

E4-TE MSA2_YEAST 39 10.08 3POJ 0.4731 0.424 0.526 0.629 0.468 0.260 0.634 0.629<br />

E5-TE BOP3_YEAST 43 9.72 3PLT 0.4791 0.426 0.539 0.632 0.439 0.259 0.651 0.615<br />

High 0.2 F1-TE ATG23_YEAST 51 5.77 2ZPN 0.4705 0.422 0.534 0.628 0.443 0.259 0.631 0.676<br />

F2-TE ATG2_YEAST 178 5.62 2DYT 0.4771 0.427 0.531 0.622 0.454 0.261 0.637 0.652<br />

F3-TE RRS1_YEAST 22 9.46 2 HBL 0.4884 0.432 0.543 0.635 0.436 0.257 0.634 0.646<br />

F4-TE Fsf1p 31 9.84 3L06 0.4936 0.438 0.579 0.643 0.414 0.255 0.635 0.649<br />

F5-TE SPS1_YEAST 55 7.55 3CZ8 0.4987 0.436 0.547 0.638 0.454 0.252 0.635 0.660<br />

0 G1-TE SHU1_YEAST 171 7.88 3R7W 0.5065 0.438 0.570 0.637 0.405 0.213 0.636 0.616<br />

G2-TE SC61G_YEAST 8 9.48 2WW9 0.5119 0.445 0.543 0.620 0.445 0.248 0.642 0.670<br />

G3-TE PDE2_YEAST 60 6.14 1MCO 0.5187 0.449 0.547 0.622 0.421 0.242 0.643 0.675<br />

G4-TE NUP53_YEAST 52 9.53 3P3D 0.6023 0.448 0.572 0.645 0.415 0.262 0.644 0.613<br />

G5-TE COX1 Protein 90 9.67 3OMA 0.6336 0.439 0.543 0.604 0.435 0.270 0.647 0.655<br />

A ; Protein hydrophobicity, B ; Salt concentration, C ; Identity number, D ; I.D. in SwissProt database, E ; Molecular weight, F ; Isoelectric point, G ; I.D in<br />

PDB databank, H ; Average surface hydrophobicity calculated by Cowan-Whittaker’s scale, I, J and K ; Average hydrophobicity calculated by the scales<br />

of Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. L and M ; Average polarity calculated by the scales of Grantham and Zimmerman,<br />

respectively. N ; Average flexibility calculated by Bhaskaran and Ponnuswamy’s scale, O ; Average bulkiness calculated by Zimmerman’s scale.<br />

52


Chapter 3<br />

Table 6. List of the identified proteins fractionated by Toyopearl-Butyl and their characterization with different properties<br />

Prot A Salt B Prot ≠ C Prot. I.D D Mr E pI F PDB G ASH H AH I AH J AH K AP L AP M AF N AB O<br />

Low 1.6 A1-TB DBF20_YEAST 65 8.68 2LAH 0.3864 0.399 0.501 0.584 0.471 0.320 0.605 0.647<br />

A2-TB YP191_YEAST 41 5.55 3C9P 0.4087 0.415 0.440 0.603 0.475 0.301 0.596 0.649<br />

A3-TB SMA1_YEAST 28 9.85 3RDJ 0.4189 0.382 0.448 0.565 0.465 0.319 0.606 0.653<br />

A4-TB YSY6_YEAST 7 10.43 3NYB 0.4192 0.388 0.450 0.588 0.560 0.279 0.583 0.655<br />

1.4 B1-TB RDS2_YEAST 50 7.87 2KOA 0.4234 0.384 0.499 0.581 0.475 0.281 0.588 0.636<br />

B2-TB PABP_YEAST 64 5.71 2X4T 0.4244 0.371 0.499 0.590 0.489 0.277 0.592 0.613<br />

B3-TB FMP46_YEAST 15 9.38 2QL8 0.4442 0.404 0.498 0.593 0.485 0.278 0.594 0.671<br />

Medium<br />

B4-TB IDHH_YEAST 47 9.24 3BLX 0.4339 0.414 0.500 0.582 0.483 0.306 0.601 0.650<br />

1.0 C1-TB TAF5_YEAST 88 7.01 3HMH 0.4447 0.374 0.502 0.595 0.483 0.306 0.590 0.620<br />

C2-TB GFD1_YEAST 21 9.76 3LCN 0.4434 0.382 0.505 0.583 0.485 0.373 0.596 0.626<br />

C3-TB YDL052C 33 9.67 3R0S 0.4412 0.384 0.499 0.588 0.475 0.312 0.575 0.668<br />

C4-TB UBP4_YEAST 105 8.50 3IRT 0.4471 0.392 0.512 0.604 0.469 0.299 0.603 0.650<br />

0.8 D1-TB Vtil1p 10 4.82 3ONJ 0.4447 0.349 0.501 0.589 0.476 0.323 0.596 0.642<br />

D2-TB LDB7_YEAST 19 9.08 2R10 0.4457 0.388 0.507 0.594 0.466 0.291 0.598 0.605<br />

D3-TB TFC4p 120 5.14 3MA5 0.4573 0.408 0.495 0.592 0.478 0.288 0.595 0.678<br />

D4-TB DRS1_YEAST 85 5.42 3RC3 0.4593 0.405 0.512 0.602 0.465 0.285 0.602 0.626<br />

D5-TB TFC1_YEAST 73 5.23 2JO4 0.4619 0.408 0.506 0.605 0.481 0.283 0.603 0.651<br />

0.6 E1-TB HIM1_YEAST 47 9.15 3P82 0.4629 0.410 0.524 0.608 0.453 0.259 0.609 0.658<br />

E2-TB TEX1_YEAST 47 8.88 3MXJ 0.4630 0.414 0.528 0.615 0.450 0.256 0.607 0.635<br />

E3-TB FAF1_YEAST 38 9.33 3R3M 0.4620 0.405 0.526 0.609 0.456 0.275 0.613 0.585<br />

E4-TB TDH3p 35 6.46 1Q32 0.4647 0.405 0.534 0.607 0.450 0.255 0.608 0.627<br />

E5-TB RPN2_YEAST 104 5.85 2X5N 0.4712 0.417 0.534 0.637 0.449 0.258 0.610 0.639<br />

High 0.2 F1-TB ICP55_YEAST 57 6.11 1VZ2 0.4644 0.414 0.527 0.607 0.456 0.273 0.615 0.641<br />

F2-TB DSD1_YEAST 47 7.53 3ANU 0.4751 0.426 0.534 0.625 0.432 0.242 0.613 0.651<br />

F3-TB PLSC_YEAST 33 9.63 1IUQ 0.4770 0.446 0.535 0.635 0.431 0.235 0.615 0.670<br />

F4-TB YMX6_YEAST 105 9.72 2JXN 0.4826 0.429 0.542 0.622 0.422 0.256 0.680 0.585<br />

F5-TB MSA2_YEAST 39 10.08 3POJ 0.4731 0.412 0.526 0.615 0.468 0.268 0.634 0.629<br />

0 G1-TB RS7A_YEAST 21 9.83 3O2Z 0.4842 0.435 0.530 0.600 0.419 0.255 0.643 0.683<br />

G2-TB YSW1_YEAST 70 6.69 3OQC 0.5018 0.437 0.547 0.625 0.417 0.255 0.629 0.648<br />

G3-TB IF4A_YEAST 44 5.02 2G9N 0.5082 0.433 0.532 0.618 0.447 0.228 0.680 0.659<br />

G4-TB SGM1_YEAST 81 4.81 2YB7 0.5210 0.448 0.572 0.633 0.412 0.248 0.639 0.638<br />

G5-TB IMP2_YEAST 19 10.01 1L0N 0.5233 0.429 0.576 0.628 0.416 0.220 0.669 0.676<br />

A ; Protein hydrophobicity, B ; Salt concentration, C ; Identity number, D ; I.D. in SwissProt database, E ; Molecular weight, F ; Isoelectric point, G ; I.D in<br />

PDB databank, H ; Average surface hydrophobicity calculated by Cowan-Whittaker’s scale, I, J and K ; Average hydrophobicity calculated by the scales<br />

of Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. L and M ; Average polarity calculated by the scales of Grantham and Zimmerman,<br />

respectively. N ; Average flexibility calculated by Bhaskaran and Ponnuswamy’s scale, O ; Average bulkiness calculated by Zimmerman’s scale.<br />

53


Chapter 3<br />

Table 7. List of the identified proteins fractionated by Toyopearl-Hexyl and their characterization with different properties<br />

Prot A Salt B Prot ≠ C Prot. I.D D Mr E pI F PDB G ASH H AH I AH J AH K AP L AP M AF N AB O<br />

Low 1.6 A1-TH POP1_YEAST 100 9.83 1R5D 0.3694 0.400 0.515 0.580 0.461 0.291 0.606 0.644<br />

A2-TH MSS51_YEAST 50 8.84 315X 0.4013 0.327 0.465 0.604 0.551 0.280 0.573 0.662<br />

A3-TH Eno1_YEAST 46 6.16 3OTR 0.4095 0.375 0.472 0.555 0.547 0.387 0.588 0.614<br />

A4-TH Eno2_YEAST 46 5.67 2XSX 0.4130 0.376 0.476 0.558 0.542 0.372 0.579 0.613<br />

1.4 B1-TH RAD16_YEAST 91 8.54 3AUX 0.4225 0.403 0.478 0.587 0.475 0.315 0.586 0.656<br />

B2-TH YAE1_YEAST 16 5.22 2XE7 0.4268 0.377 0.493 0.600 0.488 0.338 0.604 0.639<br />

B3-TH DBf2_YEAST 66 8.86 3OQ0 0.4319 0.378 0.498 0.589 0.485 0.299 0.603 0.642<br />

Medium<br />

B4-TH Pho85 cyclin 34 9.65 2PK9 0.4274 0.380 0.500 0.603 0.484 0.309 0.602 0.670<br />

1.0 C1-TH AIM37_YEAST 26 9.39 21B7 0.4444 0.395 0.507 0.603 0.475 0.305 0.592 0.676<br />

C2-TH XRN1_YEAST 175 7.04 3PIE 0.4463 0.397 0.510 0.598 0.470 0.300 0.608 0.652<br />

C3-TH MCX1_YEAST 57 7.59 3P9D 0.4471 0.403 0.516 0.608 0.470 0.308 0.619 0.675<br />

C4-TH DEBYG2_YEAST 35 6.46 3STH 0.4345 0.405 0.527 0.612 0.462 0.287 0.595 0.627<br />

0.8 D1-TH TAF5_YEAST 88 7.01 3HMH 0.4447 0.406 0.502 0.595 0.483 0.273 0.597 0.620<br />

D2-TH YO316_YEAST 07 11.66 3PLU 0.4534 0.409 0.515 0.604 0.468 0.296 0.604 0.627<br />

D3-TH UB4_YEAST 105 8.50 3IRT 0.4471 0.410 0.509 0.593 0.479 0.290 0.603 0.650<br />

D4-TH YL345_YEAST 58 6.01 2DWO 0.4536 0.414 0.520 0.611 0.464 0.295 0.594 0.659<br />

D5-TH AI2M_YEAST 97 9.67 3IFJ 0.4567 0.418 0.526 0.612 0.463 0.289 0.594 0.657<br />

0.6 E1-TH R1R1_YEAST 99 5.83 2CVY 0.4581 0.420 0.520 0.613 0.460 0.287 0.592 0.639<br />

E2-TH YE119_YEAST 14 8.49 1M94 0.4594 0.426 0.524 0.613 0.458 0.271 0.629 0.652<br />

E3-TH YJHO_YEAST 104 5.29 3SDJ 0.4612 0.427 0.526 0.609 0.450 0.277 0.632 0.667<br />

E4-TH SNZ2_YEAST 32 5.40 1DTY 0.4631 0.429 0.527 0.605 0.445 0.270 0.645 0.648<br />

E5-TH METE_YEAST 85 6.07 3L7R 0.4660 0.430 0.534 0.621 0.452 0.276 0.644 0.655<br />

High 0.2 F1-TH NIC96_YEAST 96 6.37 2RFO 0.4672 0.431 0.526 0.608 0.456 0.276 0.645 0.662<br />

F2-TH YPD1_YEAST 19 4.58 1QSP 0.4697 0.432 0.530 0.612 0.450 0.270 0.618 0.656<br />

F3-TH BOP3_YEAST 43 9.72 3PLT 0.4791 0.425 0.539 0.618 0.425 0.259 0.651 0.696<br />

F4-TH SNF12_YEAST 63 5.47 2QLV 0.4711 0.434 0.542 0.618 0.447 0.284 0.646 0.644<br />

F5-TH DYR_YEAST 24 7.67 3Q1H 0.4764 0.438 0.536 0.621 0.440 0.289 0.647 0.662<br />

0 G1-TH SPS1_YEAST 55 7.55 3CZ8 0.4987 0.442 0.527 0.626 0.415 0.252 0.630 0.649<br />

G2-TH PNG1_YEAST 42 6.22 1PNG 0.4784 0.432 0.549 0.625 0.433 0.255 0.648 0.644<br />

G3-TH PMG1_YEAST 27 8.81 4PGM 0.4843 0.435 0.552 0.627 0.427 0.248 0.649 0.657<br />

G4-TH RL20A_YEAST 20 10.30 3O58 0.4896 0.437 0.554 0.630 0.419 0.243 0.649 0.676<br />

G5-TH MLP1_YEAST 21 5.14 2XMF 0.5159 0.442 0.553 0.634 0.418 0.237 0.650 0.647<br />

A ; Protein hydrophobicity, B ; Salt concentration, C ; Identity number, D ; I.D. in SwissProt database, E ; Molecular weight, F ; Isoelectric point, G ; I.D in<br />

PDB databank, H ; Average surface hydrophobicity calculated by Cowan-Whittaker’s scale, I, J and K ; Average hydrophobicity calculated by the scales<br />

of Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. L and M ; Average polarity calculated by the scales of Grantham and Zimmerman,<br />

respectively. N ; Average flexibility calculated by Bhaskaran and Ponnuswamy’s scale, O ; Average bulkiness calculated by Zimmerman’s scale.<br />

54


Chapter 3<br />

3.1.3.3. Influence of the ligands chemistries and average protein properties<br />

3.1.3.3.1. Average hydrophobicity (AH)<br />

Average hydrophobicity is actually the average of the hydrophobicities of the overall residues<br />

in the primary structure of a protein. Average hydrophobicity has been reported for its relation<br />

with the protein structure (66). However, average hydrophobicity has been rarely investigated<br />

in the context of HIC. There are different hydrophobicity scales proposed by several research<br />

groups but there is still a debate on which one scale is the best (32). Average hydrophobicity<br />

was calculated for the tabulated proteins on the scales proposed by Cowan-Whittaker,<br />

Miyazawa-Jernigan and Tanford (Tables 5-7) (25, 92, 95). A statistically significant linear<br />

correlation (r 2 = 0.82, p < 0.0001) was reported between the average hydrophobicity and<br />

protein retention behavior (Figures 15 A-C). This confirmed that protein with high average<br />

hydrophobicity will take longer time to elute during chromatography due to the high<br />

hydrophobic interactions between proteins and adsorbents. In contrast, the protein with less<br />

hydrophobic residues in its sequence will result in less hydrophobic interactions and<br />

ultimately the protein will elute at high salt concentrations in the beginning of the<br />

chromatography. It was reported before with model proteins, that in addition to a hydrophobic<br />

surface area, it is important to have a higher overall number of hydrophobic residues in the<br />

protein sequence for a strong binding on HIC adsorbents (62). The high abundance of<br />

hydrophilic amino acids in the contact surface area was also reported to have an inverse<br />

relationship with the protein retention (113). However, these results confirmed the role of<br />

average hydrophobicity in HIC with the process proteomics approach. The hydrophobicity<br />

value for the same protein on the scale of Miyazawa-Jernigan was differed with 1 and 2 digits<br />

to the scales of Cowan-Whittaker and Tanford, respectively. The reason for this could be the<br />

different direct and indirect approaches applied by researchers while determining the<br />

hydrophobicity indices for individual amino acids (25, 92, 95). The hydrophobicity indices of<br />

55


Chapter 3<br />

amino acids by Cowan-Whittaker’s scale were derived by a direct chromatographic method<br />

and due to this it was adapted for calculating ASH. The average hydrophobicity has given a<br />

direct relationship with the retention of protein on the mentioned scales. This confirmed the<br />

role of average hydrophobicity in the retention mechanism of HIC. The global averages of the<br />

r 2 values were 0.85, 0.82 and 0.79 for the scales of Miyazawa-Jernigan, Cowan-Whittaker and<br />

Tanford, respectively. The scales of the Cowan-Whittaker and Miyazawa-Jernigan have<br />

revealed high correlation coefficient values than Tanford. It was also reported before that the<br />

scales of Cowan-Whittaker and Miyazawa-Jernigan have provided the best correlation for the<br />

prediction purposes in comparison to other scales (114).<br />

Figure 15 A: Represents the average hydrophobicity calculated by the scale of Cowan-<br />

Whittaker. The protein average hydrophobicity values of the three ligands were correlated with<br />

the respective salt concentrations where the fractions were collected. The error bars represent<br />

the standard deviations of the average hydrophobicity values in each fraction. The statistical<br />

analysis revealed a significant correlation (r 2 = 0.82; p < 0.0001).<br />

56


Chapter 3<br />

Figure 15 B: Represents the average hydrophobicity calculated by the scale of Miyazawa-<br />

Jernigan. The protein average hydrophobicity values of the three ligands were correlated with<br />

the respective salt concentrations where the fractions were collected. The error bars represent<br />

the standard deviations of the average hydrophobicity values in each fraction. The statistical<br />

analysis revealed a significant correlation (r 2 = 0.85; p < 0.0001).<br />

57


Chapter 3<br />

Figure 15 C: Represents the average hydrophobicity calculated by the scale of Tanford. The<br />

protein average hydrophobicity values of the three ligands were correlated with the respective<br />

salt concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average hydrophobicity values in each fraction. The statistical analysis<br />

revealed a significant correlation (r 2 = 0.79; p < 0.0001).<br />

Some proteins were observed to have similar values of average hydrophobicity but differed<br />

with the average surface hydrophobicity (ASH). In this case, the protein with high ASH was<br />

retained longer than the one with less ASH, although they have the same average<br />

hydrophobicity. This revealed that ASH has a more decisive role in HIC than average<br />

hydrophobicity. Although ASH is the most precise method, but the trends observed in case of<br />

average hydrophobicity cannot be ignored. The reason for this could be that, although amino<br />

acids on the surface are deciding for retention, the unfolding of proteins may occur during<br />

chromatography experiments resulting in the retention based on the total hydrophobicity of a<br />

58


Chapter 3<br />

protein. It has been reported before that the retention of proteins in reverse phase<br />

chromatography (RPC) was dependent on the total hydrophobicity of a protein (70). This can<br />

give an additional advantage to average hydrophobicity and this parameter should also be<br />

explored in RPC. The trend lines of the three ligands were statistically non significant which<br />

revealed that average hydrophobicity have no ability to differentiate among the hydrophobic<br />

characters of the three ligands (Figure 15 A-C). The reason could be that average<br />

hydrophobicity was representing the total residues of a protein instead of only surface<br />

hydrophobic residues.<br />

3.1.3.3.2. Average polarity (AP)<br />

The average polarity of the proteins has been reported in the previous studies for its<br />

contribution in the three dimensional structure of a protein; however average polarity has been<br />

rarely explored in HIC (64, 115-117). The average polarity was calculated for the tabulated<br />

proteins utilizing the scales proposed by Grantham and Zimmerman (Tables 5-7) (96, 97).<br />

The polarity values for early and late eluted proteins were more than 0.49 (49%) and less than<br />

0.41 (41%), respectively on Grantham’s scale. The same trend was also observed on<br />

Zimmerman’s scale where early eluted proteins had a higher polarity 0.35 (> 35%) than the<br />

late eluted ones 0.24 (< 24%). The membrane proteins were reported to have less polarity (<<br />

40%) while soluble proteins have higher polarity, approximately 47%. The membrane<br />

proteins with the less polarity need detergents to be separated from respective membranes due<br />

to hydrophobic interactions. (117). It was also reported before with model proteins that a<br />

decrease in the polar surface area of the protein will increase the binding affinity on the<br />

adsorbents (62). However, the effect of polarity was examined for the first time in this work<br />

with the process proteomics approach. The average polarity of the proteins has revealed a<br />

statistically significant (r 2 = 0.82, p < 0.0001) inverse correlation with the protein retention<br />

during HIC (Figure 16 A-B).<br />

59


Chapter 3<br />

Figure 16 A: Represents the average polarity calculated by Grantham’s scale. The protein<br />

average polarity values of the three ligands were correlated with the respective salt<br />

concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average polarity values in each fraction. The statistical analysis revealed a<br />

significant correlation (r 2 = 0.89; p < 0.0001).<br />

60


Chapter 3<br />

Figure 16 B: Represents the average polarity calculated by Zimmerman’s scale. The protein<br />

average polarity values of the three ligands were correlated with the respective salt<br />

concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average polarity values in each fraction. The statistical analysis revealed a<br />

significant correlation (r 2 = 0.76; p < 0.0001).<br />

This revealed that proteins with high polar residues in the protein sequence were eluted at<br />

high salt concentrations at the beginning of the chromatography. In contrast, the proteins with<br />

less polar residues in the protein sequence were eluted at the end of the chromatography at<br />

low salt concentrations. The polar residues are usually the acidic and basic amino acids (Asp,<br />

Asn, Glu, Lys, Arg, Gln). These amino acid residues are highly polar and occur on the surface<br />

of protein at an aqueous interface. The non-polar residues such as (Ala, Val, Leu, Ile, Cys,<br />

Met, Pro, Phe, and Trp) are usually found in the interior portion of the proteins (117). Based<br />

on the abundance of these residues in a protein sequence, the polarity value has been assigned.<br />

61


Chapter 3<br />

The polarity value for the same protein on Grantham’s scale was almost one digit higher than<br />

Zimmerman’s scale due to the different approaches adopted by them. The r 2 values of the<br />

polarity with the protein retention were 0.89 and 0.76 for the scales of Grantham and<br />

Zimmerman, respectively. An inverse relationship between polarity and retention volume<br />

proved that proteins with less polarity will have a more non polar surface area and will result<br />

in stronger binding to HIC surface. In contrast, proteins with high polarity will have more<br />

polar and less non polar surface area, resulting in less binding to hydrophobic adsorbents. In<br />

other words, polarity has an inverse relationship with protein hydrophobicity. An inverse<br />

relationship was also observed between polarity and flexibility; the proteins with high polarity<br />

were less flexible and eluted at high salt concentrations. Although polarity has no decisive<br />

role in retention but still it can be considered as a contributing parameter during HIC. The<br />

average polarity has shown a correlation with the retention behavior of protein, but it has no<br />

potential to differentiate among the ligands. The reason can be that this property was derived<br />

from the primary structure of a protein instead of the exposed surface residues in the three<br />

dimensional structure. HIC has more relevance with the surface residues of a protein and is<br />

rarely affected by the total residues. Instead, RPC is a chromatographic method based on the<br />

total hydrophobic residues of a protein. Due to this reason, the average polarity might have<br />

the potential to differentiate the hydrophobic characters of the reverse phase chromatographic<br />

adsorbents.<br />

3.1.3.3.3. Average flexibility (AF)<br />

The average flexibility was calculated for the tabulated proteins using the Bhaskaran and<br />

Ponnuswamy’s scale (Tables 5-7) (98). The statistical analysis revealed a direct significant<br />

correlation (r 2 = 0.80, p < 0.0001) of average flexibility with the retention behavior of the<br />

proteins during chromatography (Figure 17). The direct relationship of flexibility with<br />

protein retention could be related to the abundance of cystein residues in the primary structure<br />

62


Chapter 3<br />

of a protein forming disulfide bridges. The protein with many cystein residues / or disulfide<br />

bridges will be less flexible towards conformational changes. In contrast, the protein with<br />

high flexibility will be more flexible towards conformational changes and when it come in<br />

contact of HIC surface, it will unfold or expose the internal surface in order to increase<br />

hydrophobic interactions with the adsorbent, consequently resulting in longer retention during<br />

chromatography (33, 62). The role of average flexibility in chromatographic separation has<br />

been confirmed here for the first time with the process proteomics approach. The direct effect<br />

of flexibility on the retention behavior has also revealed the direct relationship of flexibility<br />

with proteins hydrophobicity. If a protein has high hydrophobic residues in its sequence, the<br />

higher will be the flexibility and compressibility of that protein, which is in agreement of the<br />

previous paper (119). The proteins eluted at high salt concentrations at the beginning of the<br />

chromatography have less flexible nature than those proteins eluted at the end of the<br />

chromatography. The less flexible proteins have compact structures and hydrophobic parts of<br />

the proteins exist in the inner core and resulted in less hydrophobic interactions and earlier<br />

elution during chromatography. In contrast, the highly flexible proteins will unfold during<br />

chromatography and a hydrophobic part will come to the surface which will ensue high<br />

hydrophobic interactions and resulted in later elution during chromatography. Some proteins<br />

such as C4-TH and E4-TB were eluted earlier than D1-TH and F1-TB, respectively although<br />

they have almost identical ASH. This could be due to the higher flexibilities of the latter<br />

proteins resulting in longer retention during chromatography. The same behavior was also<br />

observed with model proteins where ribonuclease S has shown higher retention on Butyl-<br />

Sepharose than other proteins, although they have identical surface hydrophobicity. It was<br />

claimed that the higher flexibility of ribonulcease S might be favoring longer retention (63).<br />

63


Chapter 3<br />

Figure 17: Represents the average flexibility calculated by Bhaskaran and Ponnuswamy’s scale.<br />

The protein average flexibility values of the three ligands were correlated with the respective<br />

salt concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average flexibility values in each fraction. The statistical analysis revealed a<br />

significant correlation (r 2 = 0.80; p < 0.0001).<br />

The flexibility and stability of the proteins were also reported for their applications in nature,<br />

where the organisms stay alive at -50°C to over 100°C. All the organisms have the same 20<br />

amino acids; however it could be the combine contribution of the amino acid residues<br />

balancing the stability and flexibility of proteins and allowing the organism to stay on both<br />

temperature extremes (120). The average flexibility has shown no relation to the molecular<br />

weight and isoelectric point values of the proteins. In other words, the flexibility values of the<br />

tabulated proteins were independent of whether a protein is larger or smaller and acidic or<br />

basic in nature. The flexibility has revealed a direct relationship with the protein retention<br />

64


Chapter 3<br />

during chromatography; however it has no ability to differentiate among the ligands for their<br />

hydrophobic characters.<br />

3.1.3.3.4. Average Bulkiness (AB)<br />

The average bulkiness was calculated for the tabulated proteins employing the scale of<br />

Zimmerman (Tables 5-7). The average bulkiness has revealed no relationship (r 2 = 0.00; p =<br />

0.86) with the retention behavior of the proteins during chromatography (Figure 18). The<br />

average bulkiness values for the tabulated proteins were in a narrow range and the reason for<br />

this could be the size exclusion effect. Although bulkiness was expected to have an effect on<br />

chromatographic behavior due to its close relationship with the size of proteins, no effect was<br />

observed in the retention behavior. In the same way, the molecular weight values of the<br />

tabulated proteins have not revealed any correlation as they were representing the small<br />

fraction of the total proteins present on each 2-D gel. It was also researched by Zimmerman<br />

who proposed the scales for bulkiness and polarity, that there was no relationship between<br />

bulkiness and polarity values of the proteins according to his scales (96). In other words, if<br />

bulkiness has no correlation with polarity then it is supposed to have no relation with protein<br />

hydrophobicity. Our results confirmed that bulkiness has no role in the chromatographic<br />

separations during HIC.<br />

65


Chapter 3<br />

Figure 18: Represents the average bulkiness calculated by Zimmerman’s scale. The protein<br />

average bulkiness values of the three ligands were correlated with the respective salt<br />

concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average bulkiness values in each fraction. The statistical analysis revealed a<br />

non significant correlation (r 2 = 0.00; p = 0.86).<br />

3.1.3.3.5. Average surface hydrophobicity (ASH)<br />

The procedure as described by J.C. Salgado et al was used to compute the ASH values for the<br />

tabulated proteins utilizing the Cowan-Whittaker’s hydrophobicity scale (Tables 5-7) (32, 92).<br />

According to this scale, the hydrophobicity values were assigned to each amino acid residue<br />

based on the amino acid retention time in RP-HPLC. A statistically significant (r 2 = 90; p<br />

value < 0.006) direct relationship was reported between ASH and retention behavior on the<br />

respective adsorbents (Figure 19). The proteins with less ASH have less hydrophobic surface<br />

residues and hence eluted at the beginning of the chromatography in comparison to the<br />

66


Chapter 3<br />

proteins with medium or high ASH values. The proteins were classified into three groups such<br />

as low, medium or highly hydrophobic, depending on the ASH and the corresponding salt<br />

concentrations on which the proteins were eluted during chromatography. The proteins eluted<br />

at high salt concentrations in the beginning of the chromatography were grouped as less<br />

hydrophobic than those proteins eluted at the middle and / or at the end of the<br />

chromatography.<br />

Figure 19: Represents the average surface hydrophobicity calculated by Cowan-Whittaker’s<br />

scale. The protein average surface hydrophobicity values of the three ligands were correlated<br />

with the respective salt concentrations where the fractions were collected. The error bars<br />

represent the standard deviations of the average surface hydrophobicity values in each<br />

fraction. The statistical analysis on average revealed a significant correlation (r 2 = 0.90; p <<br />

0.006).<br />

The ASH values were 0.413, 0.458 and 0.522 for the proteins eluted at the beginning, middle<br />

and at the end of the chromatography, respectively. These ranges were in agreement of the<br />

67


Chapter 3<br />

published papers with model proteins (32, 62). This revealed that a protein with less ASH<br />

value retained for a shorter space of time than the protein with high ASH value. This kind of<br />

behavior was also reported by several research groups that binding affinity increased with the<br />

exposed hydrophobic residues of a protein. These results confirmed that hydrophobicity of the<br />

adsorbents and proteins have the combine contribution to increase the entropy of the system<br />

and resulted in hydrophobic interactions (118). The correlation coefficients (r 2 ) values of ASH<br />

with the respective salt concentrations were 0.95, 0.94 and 0.80 for Toyopearl-Hexyl-, -Butyl<br />

and -Ether, respectively. The r 2 values in the case of Toyopearl-Ether- was very less than -<br />

Butyl and -Hexyl. The reason can be the elution of the highly hydrophobic proteins at the end<br />

of the chromatography in the case of Toyopearl-Ether which exerted an effect on the overall<br />

trend line. Some proteins were reported to have high or similar ASH values than others but<br />

retained less i.e. C3-TH and E4-TB were eluted earlier than D3-TH and F1-TB, respectively.<br />

This revealed that higher the molecular mass of the protein, the longer it will stay on the<br />

adsorbent. The behavior based on protein molecular mass was not reported for Toyopearl-<br />

Ether due to its less hydrophobic nature which resulted in less ability to entrap larger proteins.<br />

Although ASH has been reported in the prediction of retention time of model proteins,<br />

however this parameter has been never studied in HIC as a function of the cell proteome (26,<br />

87). The salt and hydrophobicity ranges were defined for the first time in this work and can be<br />

used for the separation of target proteins employing HIC. ASH has revealed a direct<br />

relationship with the retention of proteins; however it has a limited ability to differentiate<br />

among different ligands.<br />

3.1.3.4. Comparative studies of the hydrophobic ligands based on average protein<br />

properties<br />

There are several reports where various ligands have been compared for their hydrophobicity<br />

utilizing the retention time of model proteins (22, 33, 61, 62, 99). However, the<br />

hydrophobicity of the ligands has been never explored with a cell proteome. The<br />

68


Chapter 3<br />

hydrophobicity of the ligands was investigated in this work with the yeast cell proteome.<br />

However, the comparison of ligands has not fully differentiated with the proteome wide<br />

approach (Figure 19). The reason could be the involvement of several protein properties<br />

affecting the retention behavior in HIC. Several protein properties have been demonstrated for<br />

their correlation with protein retention and their ability to differentiate among the<br />

hydrophobic ligands. The protein properties such as average surface hydrophobicity, average<br />

hydrophobicity, average flexibility and average polarity have shown a correlation with the<br />

chromatographic separation of the proteins. However, these properties may have less ability<br />

to differentiate among the ligands (Table 8).<br />

Table 8: Protein properties and their ability to differentiate among ligands and correlation with<br />

chromatographic behavior during HIC<br />

Protein properties<br />

Differentiation Correlation<br />

Average hydrophobicity × +<br />

Average polarity × -<br />

Average flexibility × +<br />

Average bulkiness × ×<br />

Average surface hydrophobicity Trend +<br />

+; indicates positive correlation, - ; indicates negative correlation; ×; indicates no correlation<br />

The results revealed that ligands cannot be fully differentiated based on protein properties and<br />

can only be estimated based on retention time of the model proteins. The reason could be that<br />

several proteins are supposed to be eluted at specific retention time during chromatographic<br />

separation of a cell proteome as a function of ligand chemistry. Another observation was that<br />

69


Chapter 3<br />

ligands revealed less difference at high salt concentrations than at low salt concentrations. In<br />

other words, the differences among the ligands were less at high hydrophobic conditions than<br />

at low hydrophobic conditions. This kind of behavior was also reported in our group with<br />

model proteins, where the differences among the adsorbents were higher at low salt<br />

concentrations than at the high salt concentrations (121). The reason for this could be that at<br />

the extreme hydrophobic conditions, the hydrophobic potential of the mobile phase has<br />

covered the potential of the ligands to demonstrate according to their hydrophobicities. At low<br />

salt concentrations, the mobile phase has low hydrophobicity and the adsorption was more<br />

dependent on the hydrophobicity of the ligands.<br />

3.1.3.5. Application note<br />

Several important proteins have been separated through chromatographic and electrophoretic<br />

separations and identified through MALDI-ToF-MS. This work can help those interested in<br />

the purification of proteins listed in Tables 5-7, with the methodology as described in chapter<br />

2. It will simplify the work of researchers and they will not have to scout for chromatographic<br />

experiments to figure out the elution position of any tabulated protein. Any of the target<br />

protein can be purified by isocratic chromatography, without going through trial and error<br />

chromatographies. Some of the important enzymes were Isocitrate dehydrogenase (eluted at<br />

1.4 M salt gradient on Toyopearl-Butyl), Phosphoglycerate mutase 1 (eluted at 0 M on<br />

Toyopearl-Hexyl), Triosephosphate isomerase (1.4 M on Toyopearl-Ether), Methyl<br />

transferase (0.6 M on Toyopearl-Hexyl), Enolase 1 and 2 (1.6 M on Toyopearl-Hexyl), D-<br />

serine dehydratase (0.2 M on Toyopearl-Butyl), Exoribonuclease 1 and Glyceraldehyde 3<br />

phosphate dehydrogenase (1.0 M on Toyopearl-Hexyl).<br />

70


Chapter 3<br />

3.1.4. Partial conclusions<br />

• The hydrophobicity of the three ligands has been investigated for the first time with<br />

the proteome wide approach utilizing different protein properties.<br />

• The ligands were not fully differentiated with the proteome wide approach, but a trend<br />

was observed.<br />

• Several protein properties such as average hydrophobicity, average flexibility, average<br />

polarity, average bulkiness and average surface hydrophobicity were investigated for<br />

their correlation with the protein retention during chromatography. Average polarity<br />

has shown an inverse relationship to the retention behavior of the proteins during HIC.<br />

Average surface hydrophobicity, average hydrophobicity and average flexibility have<br />

revealed direct relationship with the protein retention during chromatography.<br />

• Among all the properties, ASH was the only property which has a decisive role in HIC.<br />

• The large scale identification and characterization of the proteins has explored the<br />

contaminant profile of the yeast cell proteome and is an addition to the proteomics<br />

technology.<br />

• The data obtained by exploring the chromatographic behavior of the yeast cell<br />

proteome on the three different ligands can be used to design the purification models<br />

of recombinant proteins utilizing the in silico approach.<br />

71


Chapter 3<br />

3.2. Influence of the different base support chemistries during<br />

hydrophobic interaction chromatography<br />

3.2.1. Summary<br />

The comparison of the different base support chemistries as a function of the yeast cell<br />

proteome was explored for the first time in this work. This work was planned in the<br />

continuation of a previous report from our group where several ion exchange beads were used<br />

to investigate the separation behavior of the insect cell proteome instead of model proteins<br />

(19). Although several conclusions were made in that work, but still the approach was not<br />

complete in its essence and several techniques were missing. That approach was further<br />

progressed by exploring the different base supports in HIC employing some additional<br />

technologies. In this study, the Saccharomyces cerevisiae was selected as a host cell proteome<br />

due to the reasons such as its availability in protein data bank and it has the advantage of<br />

providing post translational modifications. An experimentally demanding and laborious<br />

approach was adopted ranging from chromatographic and electrophoretic separation of the<br />

cell proteome to identification and analysis of the proteins in each chromatographic fraction<br />

by mass spectrometry and bioinformatics tools. Moreover, certain protein properties were<br />

examined for their correlation to the chromatographic separation during HIC. These<br />

properties were also investigated for their ability to differentiate among the different base<br />

supports. In addition to the role of base supports chemistries, this work was planned to<br />

produce the data with the process proteomics approach which could be used to design the<br />

purification models for recombinant proteins utilizing the in silico approach.<br />

72


Chapter 3<br />

3.2.2. Chemistry of the base supports<br />

The adsorbents such as Toyopearl-Phenyl, Sepharose-Phenyl and Source-Phenyl were used<br />

with different base supports harbored phenyl as a common ligand. The physical properties of<br />

these adsorbents have been given in Table 9 as provided by the Vendor Company (99). The<br />

commercial adsorbents are usually composed of ligands and base supports of various nature.<br />

There are several base supports usually utilized in HIC such as Sepaharose ® , Toyopearl ® and<br />

Source ®<br />

that are composed of certain polymers in backbone chemistries i.e agarose,<br />

methacrylate and polystyrene. These polymers are usually of hydrophobic nature but were<br />

modified to the hydrophilic nature by coupling to certain ether and hydroxyl groups (99). The<br />

same was the case with polystyrene based adsorbents, however its chemistry has not reported<br />

yet. The chemistry of the three mentioned adsorbents has been shown in Table 10.<br />

Table 9: Physical properties of the adsorbents with different base support (99)<br />

Adsorbents Base supports Ligands Bead size<br />

(µm)<br />

Pore size<br />

(nm)<br />

a DBC<br />

(mg/mL)<br />

Toyopearl-<br />

Phenyl<br />

Sepharose-<br />

Phenyl<br />

Toyopearl ® Phenyl 100 100 40<br />

Sepharose ® Phenyl 34 65 35<br />

Source-Phenyl Source ® Phenyl 15 34 25<br />

a DBC; Dynamic binding capacity<br />

73


Chapter 3<br />

Table 10: The chemistry of base supports and ligands of the three adsorbents<br />

Base supports Ligand Adsorbents<br />

Polymethacrylate (Toyopearl®)<br />

Phenyl Ligand<br />

Toyopearl-Phenyl<br />

H CH 3 2<br />

C C<br />

O C<br />

O<br />

(HO) R<br />

H O 2<br />

C<br />

O<br />

C<br />

H 2<br />

C<br />

C<br />

CH 3<br />

H 2<br />

C<br />

C<br />

C O<br />

O<br />

R (OH)<br />

O<br />

C O<br />

C<br />

CH 3<br />

O<br />

HW-<br />

65<br />

O<br />

Agarose (Sepharose®)<br />

Phenyl Ligand<br />

Sepharose-Phenyl<br />

CH 2 OH<br />

OH O<br />

H<br />

H<br />

H<br />

O OH<br />

H<br />

O<br />

H<br />

O<br />

H OH<br />

O<br />

H<br />

CH 2<br />

H<br />

H<br />

O<br />

PS<br />

O<br />

H 2<br />

C<br />

OH<br />

C<br />

H<br />

H 2<br />

C<br />

O<br />

Polystyrene divinyl benzene (Source®)<br />

Phenyl Ligand<br />

Source-Phenyl<br />

O<br />

Not reported yet<br />

74


Chapter 3<br />

3.2.3. Results and Discussion<br />

3.2.3.1. Chromatographic fractionation<br />

Scouting experiments were performed to explore the most suitable operational conditions for<br />

the chromatographic separations of the yeast cell proteome. These standardized<br />

chromatographic conditions are summarized in Table 11. A decreasing gradient of<br />

ammonium sulfate was performed from 1.6 to 0 M and seven fractions were collected at<br />

specific salt concentrations in order to investigate the proteins available in each<br />

chromatographic fraction. The chromatograms in Figure 20 depict the resolution behavior of<br />

the three base supports for the same crude extract. There was no clear matching in the<br />

separation profiles of the three adsorbents due to the differences in their base support<br />

chemistries. The peaks with the number of 6, 4 and 1 correspond to Sepharose-Phenyl,<br />

Source-Phenyl and Toyopearl-Phenyl, respectively. A high resolution was observed in case of<br />

Sepharose-Phenyl and Source-Phenyl, due to their smaller bead size than Toyopearl-Phenyl.<br />

One large peak was observed for Toyopearl-Phenyl, the reason could be the larger bead size<br />

of the base support, which resulted in a low resolution during chromatographic separation.<br />

The overall resolution of the three chromatograms was not high. The reason could be the high<br />

amount of the sample load applied during chromatographic fractionation. In HIC, it is the<br />

initial effort to report the selectivity of phenyl media for a crude extract. The experimental<br />

approach applied in this work, where fractionation of a soluble cell proteome was done<br />

instead of model proteins will pave the way for a direct comparison between adsorbents in<br />

terms of hydrophobicity towards proteins derived from a complex biological mixture.<br />

Information of this kind of work has been rarely available in literature where they utilized a<br />

host cell proteome for chromatographic fractionation instead of the model proteins (63).<br />

A high amount of proteins (almost 50%) was lost in the flow through and the remaining<br />

proteins were retained in the column. This kind of behavior was also reported in Section 3.1,<br />

75


Chapter 3<br />

where approximately 50% of proteins were washed out during flow through. This revealed the<br />

high abundance of less hydrophobic proteins in the yeast cell proteome in comparison to the<br />

highly hydrophobic proteins. This also confirmed the less hydrophobic nature of the yeast cell<br />

proteome (104).<br />

Table 11: Standardized chromatographic conditions<br />

Feature<br />

Description<br />

Column volume<br />

5 mL<br />

Flow rate<br />

1 mL / min<br />

Equilibration buffer 10 CV a , 1.6 M, (NH 4 ) 2 SO 4 pH 7.5<br />

Sample load<br />

5 CV a , (8 mg / mL)<br />

Linear Gradient<br />

25 CV a<br />

Fraction volume<br />

3 mL<br />

CV a Column volume<br />

76


Chapter 3<br />

Figure 20: Represent the chromatographic profiles obtained <strong>after</strong> the proteome fractionation by HIC using Sepharose-Phenyl, Toyopearl-Phenyl and<br />

Source-Phenyl. Proteins were loaded in a mobile phase composed of equilibration buffer 1.6 M (NH 4 ) 2 SO 4 at pH 7.5. The elution was exerted for 25 CV.<br />

The fractions were collected in seven groups utilizing the following salt concentrations as cut off values: 1 (1.6 M), 2 (1.4 M), 3 (1.0 M), 4 (0.8 M), 5 (0.6<br />

M), 6 (0.2 M) and 7 (0.0 M). Each fraction was collected by pooling in triplicate. (- Refers to absorbance 280 nm, --- conductivity mS / cm).<br />

77


Chapter 3<br />

3.2.3.2. The electrophoretic separation and proteins identification<br />

The chromatographic fractions were further explored for their protein content by 2-D PAGE.<br />

The protein concentrations of the chromatographic fractions were kept similar during their<br />

electrophoretic separation by 2-D PAGE. This would give an insight to the total mass and the<br />

number of protein contaminants recovered in each chromatographic fraction (19). The 2-D<br />

gels were important for mapping the chromatographic separation of the yeast cell proteome<br />

and to excise certain spots for protein identifications (Figures 21-23). The seventy seven<br />

proteins were identified by MALDI-ToF-MS from all the gels representing specific<br />

chromatographic fractions (Tables 12-14). The four proteins were visible on the gels with the<br />

naked eyes, however were not visible when photographed and it is sometimes the case with<br />

the weakly stained proteins (105-111). The proteins identified from each chromatographic<br />

fraction were collected at different salt concentrations and can be separated through isocratic<br />

chromatography. This will avoid trial and error chromatographies to figure out the elution<br />

position of any of the tabulated protein. The large scale identification and characterization of<br />

the proteins has explored the contaminant profile of the yeast cell proteome and is an addition<br />

to the proteomics technology. The identified proteins and the corresponding 2-D gels can also<br />

be used in future to identify proteins by relying on 2-D gel electrophoresis image databases<br />

(123). The previous work on HIC was limited to a few homologous model proteins and<br />

average surface hydrophobicity parameter, which may not portray a global picture of the<br />

retention mechanism during HIC (23, 62). However, the global analysis of the base supports<br />

and protein properties as a function of complex cell proteome has overwhelmed the<br />

shortcomings in the previous approaches.<br />

78


Chapter 3<br />

Figure 21: Represent 2-D D gels performed for the samples collected by Toyopearl-Phenyl. Each<br />

gel represents the fractions collected at different salt concentrations such as A (1.6 M), B (1.4<br />

M), C (1 M), D (0.8 M), E (0.6 M), F (0.2 M) and G (0.0 M). The spots indicated by arrows were<br />

selected for identifications.<br />

79


Chapter 3<br />

Figure 22: Represent 2-D D gels performed for the samples fractionated by Sepharose-Phenyl.<br />

Each gel represents the fractions collected at different salt concentrations such as A (1.6 M), B<br />

(1.4 M), C (1 M), D (0.8 M), E (0.6 M), F (0.2 M) and G (0.0 M). The spots indicated icated by arrows were<br />

selected for identifications. . The encircled spots have been explained in section 3.2.3.2.<br />

80


Chapter 3<br />

Figure 23: Represent 2-D D gels performed for the samples fractionated by Source-Phenyl. Each<br />

gel represents the fractions collected at different salt concentrations such as A (1.6 M), B (1.4<br />

M), C (1 M), D (0.8 M), E (0.6 M), F (0.2 M), G (0.0 M). The spots indicated by arrows were<br />

selected for identifications. . The encircled spots have been explained in section 3.2.3.2.<br />

81


Chapter 3<br />

Table 12. List of the identified proteins fractionated by Toyopearl-Phenyl and their characterization with different properties<br />

Prot A Salt B Prot ≠ C Prot. I.D D Mr E pI F PDB G ASH H AH I AH J AH K AP L AP M AF N AB O<br />

Low 1.6 A1-TP KIN4_YEAST 90 9.50 2LAH 0.3864 0.356 0.486 0.570 0.495 0.303 0.540 0.622<br />

Medium<br />

A2-TP SLU7P 41 8.57 2HBP 0.3901 0.319 0.419 0.604 0.464 0.365 0.591 0.613<br />

A3-TP YDR532cp 44 4.80 1RW7 0.4015 0.375 0.488 0.605 0.496 0.326 0.611 0.668<br />

A4-TP ENO2_YEAST 46 5.67 2XHO 0.4198 0.401 0.467 0.608 0.494 0.298 0.579 0.613<br />

1.4 B1-TP HIBCH_YEAST 56 8.57 3BPT 0.4351 0.398 0.496 0.609 0.509 0.321 0.584 0.651<br />

B2-TP YKL215Cp 49 8.11 1K1D 0.4375 0.403 0.507 0.611 0.526 0.293 0.587 0.597<br />

B3-TP PMG1_YEAST 27 8.81 3R7A 0.4427 0.385 0.520 0.608 0.467 0.297 0.608 0.644<br />

B4-TP DENR_YEAST 22 8.78 2WPV 0.4449 0.394 0.523 0.609 0.473 0.295 0.615 0.644<br />

1.0 C1-TP UPT 28 6.75 2EHJ 0.4451 0.408 0.528 0.606 0.467 0.288 0.607 0.669<br />

C2-TP SPC2_YEAST 20 9.22 3PSI 0.4445 0.415 0.539 0.612 0.485 0.285 0.573 0.684<br />

C3-TP SEC4_YEAST 23 6.62 1G16 0.4471 0.411 0.522 0.607 0.465 0.286 0.604 0.624<br />

C4-TP PCL2_YEAST 35 9.72 2PMI 0.4478 0.400 0.523 0.613 0.464 0.288 0.616 0.668<br />

0.8 D1-TP PES4_YEAST 70 9.48 2V6Z 0.4609 0.404 0.531 0.622 0.478 0.278 0.613 0.648<br />

D2-TP PGM2_YEAST 63 6.18 3OLP 0.4624 0.404 0.532 0.621 0.453 0.264 0.612 0.625<br />

D3-TP YKL084W 13 6.42 3QWP 0.4672 0.418 0.535 0.618 0.456 0.280 0.625 0.659<br />

0.6 E1-TP ADH1_YEAST 36 6.21 3MEQ 0.4643 0.426 0.560 0.633 0.457 0.267 0.639 0.604<br />

E2-TP LOC1_YEAST 23 10.31 3O2Z 0.4842 0.429 0.542 0.624 0.464 0.252 0.629 0.634<br />

E3-TP DSD1_YEAST 47 7.53 3GWQ 0.4857 0.426 0.547 0.625 0.462 0.242 0.644 0.611<br />

E4-TP YPT52_YEAST 26 4.49 3RSB 0.4713 0.428 0.565 0.630 0.453 0.265 0.628 0.621<br />

High 0.2 F1-TP RTG2_YEAST 65 8.17 3OYC 0.4871 0.422 0.536 0.625 0.440 0.262 0.617 0.637<br />

F2-TP VATE_YEAST 26 5.33 2KZ9 0.4843 0.438 0.578 0.629 0.455 0.270 0.628 0.658<br />

F3-TP IDHH_YEAST 47 9.29 2RNG 0.4938 0.434 0.542 0.634 0.454 0.256 0.641 0.650<br />

F4-TP COQ3_YEAST 36 6.51 2YBO 0.5010 0.441 0.533 0.630 0.445 0.270 0.645 0.664<br />

0 G1-TP LONM_YEAST 127 5.43 2X36 0.5109 0.447 0.625 0.640 0.440 0.264 0.649 0.639<br />

G2-TP AIM4_YEAST 14 10.31 3RLS 0.5120 0.445 0.612 0.638 0.442 0.253 0.679 0.613<br />

G3-TP SPS1_YEAST 55 7.55 3CM1 0.5340 0.451 0.585 0.656 0.444 0.242 0.674 0.660<br />

G4-TP RL16B_YEAST 22 10.54 3R3T 0.5488 0.489 0.619 0.638 0.432 0.241 0.681 0.649<br />

A ; Protein hydrophobicity, B ; Salt concentration, C ; Identity number, D ; I.D. in SwissProt database, E ; Molecular weight, F ; Isoelectric point, G ; I.D in<br />

PDB databank, H ; Average surface hydrophobicity calculated by Cowan-Whittaker’s scale, I, J and K ; Average hydrophobicity calculated by the scales<br />

of Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. L and M ; Average polarity calculated by the scales of Grantham and Zimmerman,<br />

respectively. N ; Average flexibility calculated by Bhaskaran and Ponnuswamy’s scale, O ; Average bulkiness calculated by Zimmerman’s scale.<br />

82


Chapter 3<br />

Table 13. List of the identified proteins fractionated by Sepharose-Phenyl and their characterization with different properties.<br />

Prot A Salt B Prot≠ C Prot. I.D D Mr E pI F PDB G ASH H AH I AH J AH K AP L AP M AF N AB O<br />

Low 1.6 A1-PS KIN4_YEAST 90 9.50 2LAH 0.3864 0.356 0.486 0.570 0.495 0.303 0.540 0.622<br />

Medium<br />

A2-PS ORF194 21 8.50 2X5T 0.4184 0.384 0.497 0.594 0.490 0.321 0.604 0.628<br />

A3-PS YJM789 93 6.05 3A15 0.4245 0.334 0.448 0.566 0.532 0.343 0.579 0.604<br />

A4-PS RPIA_YEAST 28 5.63 3KWM 0.4351 0.408 0.526 0.625 0.465 0.288 0.584 0.630<br />

1.4 B1-PS YHR042Wp 33 8.45 2XNC 0.4194 0.366 0.462 0.559 0.507 0.348 0.602 0.648<br />

B2-PS NDUS6_YEAST 12 9.37 3RHD 0.4202 0.405 0.520 0.600 0.462 0.289 0.608 0.605<br />

B3-PS BIG1_YEAST 39 5.20 3LTL 0.4231 0.368 0.542 0.586 0.440 0.326 0.610 0.694<br />

B4-PS LEUC_YEAST 85 5.61 3Q68 0.4342 0.401 0.492 0.578 0.465 0.285 0.612 0.626<br />

1.0 C1-PS ICP55_YEAST 57 6.11 2XPG 0.4281 0.383 0.495 0.607 0.456 0.273 0.615 0.641<br />

C2-PS PRS7_YEAST 51 5.32 3BPT 0.4351 0.383 0.496 0.594 0.487 0.324 0.612 0.623<br />

C3-PS MSC3_YEAST 80 9.14 2WGP 0.4522 0.401 0.518 0.579 0.517 0.280 0.612 0.601<br />

C4-PS RE114_YEAST 48 9.26 2PMI 0.4478 0.377 0.506 0.578 0.481 0.253 0.629 0.655<br />

0.8 D1-PS PNPP_YEAST 34 6.00 1VJR 0.4473 0.414 0.547 0.621 0.428 0.243 0.601 0.635<br />

D2-PS CDC73_YEAST 44 7.72 3OIP 0.4499 0.415 0.509 0.599 0.476 0.284 0.620 0.641<br />

D3-PS YER145c-a-yeast 16 9.51 3HIE 0.4515 0.418 0.539 0.624 0.454 0.274 0.609 0.653<br />

0.6 E1-PS RS3A1_YEAST 28 10 2KHJ 0.4535 0.415 0.547 0.586 0.482 0.258 0.608 0.649<br />

E2-PS KPYK1_YEAST 54 8.00 1A3X 0.4583 0.411 0.537 0.624 0.449 0.264 0.616 0.645<br />

E3-PS UB4_YEAST 105 8.44 2V6Z 0.4609 0.411 0.539 0.605 0.454 0.249 0.612 0.651<br />

E4-PS VATC_YEAST 44 6.22 1U7l 0.4615 0.415 0.534 0.621 0.451 0.240 0.623 0.667<br />

High 0.2 F1-PS GPN_YEAST 39 4.43 1A4R 0.4647 0.432 0.628 0.628 0.454 0.237 0.624 0.654<br />

F2-PS METE_YEAST 85 6.07 3EG3 0.4757 0.410 0.534 0.621 0.452 0.234 0.632 0.655<br />

F3-PS ADH1_YEAST 36 6.21 3MEQ 0.4643 0.426 0.560 0.633 0.426 0.227 0.639 0.604<br />

0 G1-PS CEX1_YEAST 84 5.27 1Z3H 0.4795 0.418 0.540 0.621 0.446 0.242 0.617 0.655<br />

G2-PS DPM1_YEAST 30 7.70 2GEJ 0.4997 0.452 0.572 0.635 0.411 0.245 0.640 0.666<br />

G3-PS 54S L22 34 9.98 3R3T 0.5488 0.449 0.660 0.655 0.412 0.298 0.639 0.635<br />

G4-PS MCM3_YEAST 107 5.26 3F9V 0.5853 0.455 0.628 0.603 0.410 0.214 0.624 0.628<br />

A ; Protein hydrophobicity, B ; Salt concentration, C ; Identity number, D ; I.D. in SwissProt database, E ; Molecular weight, F ; Isoelectric point, G ; I.D in<br />

PDB databank, H ; Average surface hydrophobicity calculated by Cowan-Whittaker’s scale, I, J and K ; Average hydrophobicity calculated by the scales<br />

of Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. L and M ; Average polarity calculated by the scales of Grantham and Zimmerman,<br />

respectively. N ; Average flexibility calculated by Bhaskaran and Ponnuswamy’s scale, O ; Average bulkiness calculated by Zimmerman’s scale.<br />

83


Chapter 3<br />

Table 14. List of the identified proteins fractionated by Source-Phenyl and their characterization with different properties<br />

Prot A Salt B Prot ≠ C Prot. I.D D Mr E pI F PDB G ASH H AH I AH J AH K AP L AP M AF N AB O<br />

Low 1.6 A1-SP ENT4_YEAST 28 9.38 2QY7 0.3715 0.374 0.484 0.566 0.493 0.314 0.594 0.638<br />

Medium<br />

A2-SP SLU7P 41 8.57 2HBP 0.3901 0.334 0.419 0.604 0.464 0.365 0.591 0.613<br />

A3-SP RS6_YEAST 26 10.44 2XZM 0.4068 0.366 0.477 0.552 0.498 0.373 0.583 0.637<br />

A4-SP RM32_YEAST 21 9.86 3BYI 0.4174 0.373 0.492 0.562 0.512 0.358 0.612 0.653<br />

1.4 B1-SP YN93_YEAST 32 4.97 2PKP 0.4169 0.366 0.475 0.548 0.501 0.348 0.574 0.663<br />

B2-SP LRG1_YEAST 116 8.54 3BYI 0.4174 0.361 0.490 0.611 0.443 0.269 0.585 0.664<br />

B3-SP TDXH_AERPE 28 6.32 3A5W 0.4175 0.374 0.484 0.580 0.493 0.314 0.587 0.665<br />

B4-SP VID22_YEAST 102 5.06 3IFW 0.4232 0.385 0.488 0.619 0.446 0.314 0.603 0.635<br />

1.0 C1-SP BIG1_YEAST 39 5.20 3LTL 0.4231 0.403 0.498 0.586 0.440 0.326 0.610 0.694<br />

C2-SP VAM6_YEAST 97 6.61 1KMD 0.4255 0.386 0.524 0.564 0.452 0.302 0.589 0.674<br />

C3-SP SKG3_YEAST 114 8.37 3MP8 0.4324 0.392 0.498 0.605 0.486 0.305 0.631 0.616<br />

C4-SP Cha1p 22 9.70 3O05 0.4375 0.398 0.506 0.605 0.472 0.302 0.601 0.631<br />

0.8 D1-SP SYFA_YEAST 57 5.53 2WP1 0.4441 0.410 0.521 0.610 0.464 0.291 0.591 0.653<br />

D2-SP PWP2_YEAST 103 4.94 2Y9D 0.4442 0.415 0.518 0.617 0.464 0.275 0.604 0.635<br />

D3-SP YH139_YEAST 11 9.75 2YBA 0.4500 0.408 0.537 0.611 0.446 0.280 0.590 0.658<br />

0.6 E1-SP YKL085W 35 8.46 1MLD 0.4493 0.409 0.546 0.617 0.451 0.255 0.604 0.631<br />

E2-SP 1pk1p 11 6.96 2XAM 0.4537 0.410 0.545 0.623 0.443 0.304 0.634 0.697<br />

E3-SP MTC1_YEAST 53 4.50 3KOX 0.4576 0.415 0.572 0.619 0.418 0.269 0.657 0.651<br />

High 0.2 F1-SP YDL186Wp 32 5.46 2J04 0.4619 0.424 0.590 0.651 0.428 0.246 0.643 0.625<br />

F2-SP RDL2_YEAST 16 9.65 3L7S 0.4699 0.431 0.532 0.615 0.424 0.294 0.660 0.658<br />

0 G1-SP PAP_YEAST 64 7.95 3C66 0.4771 0.427 0.633 0.611 0.418 0.294 0.668 0.662<br />

G2-SP VPS64_YEAST 67 6.97 3R42 0.4927 0.423 0.598 0.651 0.426 0.283 0.639 0.628<br />

G3-SP BRE2 58 5.78 2QIY 0.4797 0.453 0.633 0.648 0.427 0.226 0.634 0.631<br />

G4-SP SPP41 99 8.93 2XF5 0.5263 0.464 0.617 0.653 0.410 0.219 0.651 0.629<br />

A ; Protein hydrophobicity, B ; Salt concentration, C ; Identity number, D ; I.D. in SwissProt database, E ; Molecular weight, F ; Isoelectric point, G ; I.D in<br />

PDB databank, H ; Average surface hydrophobicity calculated by Cowan-Whittaker’s scale, I, J and K ; Average hydrophobicity calculated by the scales<br />

of Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. L and M ; Average polarity calculated by the scales of Grantham and Zimmerman,<br />

respectively. N ; Average flexibility calculated by Bhaskaran and Ponnuswamy’s scale, O ; Average bulkiness calculated by Zimmerman’s scale.<br />

84


Chapter 3<br />

3.2.3.3. Influence of the base support chemistries and average protein properties<br />

3.2.3.3.1. Average hydrophobicity (AH)<br />

Average hydrophobicity is the classical parameter based on the overall hydrophobic residues<br />

of a protein. It was reported before that in addition to hydrophobic surface residues, it is<br />

important to have a higher overall number of hydrophobic residues in the protein sequence for<br />

the stronger binding to the HIC adsorbents (62). However, the average hydrophobicity has<br />

been never investigated with the process proteomics approach. Hence, the three different<br />

hydrophobicity scales were used to calculate average hydrophobicity from the primary<br />

structures of the tabulated proteins (Tables 12-14). A statistically significant (r 2 =0.86, p <<br />

0.0001) linear relationship was reported between the average hydrophobicity and protein<br />

retention behavior on different base supports (Figure 24 A-C). In some cases, the proteins<br />

were reported to have similar average hydrophobicity values but differed with average surface<br />

hydrophobicity. This revealed that two proteins can have the same amount of total<br />

hydrophobic residues in the primary sequence but can differ with surface hydrophobic<br />

residues. The average hydrophobicity values for some proteins were similar; however their<br />

elution position was different from each other. In this case, the proteins with high ASH<br />

retained longer than the one with low ASH, although they have similar average<br />

hydrophobicity. The proteins such as B3-TP, and F1-TP retained longer during<br />

chromatography than A4-TP and E3-TP, respectively. These results confirmed that ASH has a<br />

decisive role in protein retention during chromatography. The overall values of average<br />

hydrophobicity revealed that the scale of Miyazawa-Jernigan was differed with 1 and 2 digits<br />

to the scales of Cowan-Whittaker and Tanford, respectively. The reason for this difference<br />

could be the various direct and indirect approaches, while determining the hydrophobicity<br />

indices for individual amino acids (25, 92, 95). The mentioned scales were calculated with<br />

quite different approaches and often correlated very poorly (124). The global averages of the<br />

85


Chapter 3<br />

r 2 values from the data of the three base supports were 0.90, 0.89 and 0.79 for the scales of<br />

Miyazawa-Jernigan, Cowan-Whittaker and Tanford, respectively. It was also reported before<br />

that the scales of Cowan-Whittaker and Miyazawa-Jernigan have provided the best correlation<br />

for the prediction purposes in comparison to other scales (114). Although ASH is the most<br />

precise method, but the trends observed in case of average hydrophobicity cannot be ignored.<br />

The reason for this could be that, although amino acids on the surface are deciding for<br />

retention, the unfolding of proteins may occur during chromatographic experiments resulting<br />

in retention based on the total hydrophobicity of a protein.<br />

Figure 24 A: Represents the average hydrophobicity calculated by the scale of Cowan-<br />

Whittaker. The protein average hydrophobicity values of the three base supports were<br />

correlated with the respective salt concentrations where the fractions were collected. The error<br />

bars represent the standard deviations of the average hydrophobicity values in each fraction.<br />

The statistical analysis revealed a significant correlation (r 2 = 0.89; p < 0.0001).<br />

86


Chapter 3<br />

Figure 24 B: Represents the average hydrophobicity calculated by the scale of Miyazawa-<br />

Jernigan. The protein average hydrophobicity values of the three base supports were<br />

correlated with the respective salt concentrations where the fractions were collected. The error<br />

bars represent the standard deviations of the average hydrophobicity values in each fraction.<br />

The statistical analysis revealed a significant correlation (r 2 = 0.90; p < 0.0001).<br />

87


Chapter 3<br />

Figure 24 C: Represents the average hydrophobicity calculated by the scale of Tanford. The<br />

protein average hydrophobicity values of the three base supports were correlated with the<br />

respective salt concentrations where the fractions were collected. The error bars represent the<br />

standard deviations of the average hydrophobicity values in each fraction. The statistical<br />

analysis revealed a significant correlation (r 2 = 0.79; p < 0.0001).<br />

It was reported that the folded proteins have less non polar surface area in comparison to the<br />

unfolded ones (125). This confirmed that during unfolding, the hydrophobic residues of the<br />

proteins were exposed on the surface and resulted in longer retention during chromatography.<br />

It was also described in the mentioned report that there has been no relation to the non-polar<br />

surface area and the size of protein. In the same way, no relation was observed here between<br />

the size and hydrophobicity of the proteins. It has been reported that there were very few<br />

buried residues which have zero accessible surface area. Most of the protein interior was<br />

composed of amino acids have contact with the solvent (125). This also revealed that most of<br />

88


Chapter 3<br />

the amino acids in a protein had a contribution in the retention behavior and due to this reason,<br />

the correlation of average hydrophobicity with retention seem sensible. The average<br />

hydrophobicity has been also reported for its significance in reverse phase chromatography<br />

(RPC). The RPC is based on the overall hydrophobicity derived from the primary structure<br />

instead of only surface residues of a protein (126). This revealed that average hydrophobicity<br />

and average surface hydrophobicity have decisive roles in HIC and RPC, respectively. The<br />

analysis of the two methods of protein hydrophobicity confirmed that ASH was more relevant<br />

to HIC than average hydrophobcity in terms of chromatographic separations. The trend lines<br />

(Figure 24 A-C) revealed that although average hydrophobicity has a role in HIC, but it has<br />

no ability to differentiate among the hydrophobic characters of the three base supports.<br />

3.2.3.3.2. Average polarity (AP)<br />

The average polarity was for the first time investigated for its correlation with protein<br />

retention behavior in HIC as a function of the cell proteome. The average polarity was<br />

calculated for the proteins using the scales proposed by Grantham and Zimmerman (Tables<br />

12-14) (96, 97). A statistically significant (r 2 = 0.82, p < 0.0001) inverse relationship was<br />

observed between polarity and retention behavior of the proteins on different base supports.<br />

The proteins eluted at high salt concentrations were highly polar than the proteins eluted at<br />

low salt concentrations. In other words, the proteins eluted at the beginning of the<br />

chromatography were more polar than those proteins eluted at the end of the chromatography.<br />

The average polarity values of the Grantham’s scale were higher in early eluted proteins (><br />

0.49) than late eluted proteins (< 0.41). The same trend was also observed on Zimmerman’s<br />

scale, where early eluted proteins had high average polarity (> 0.35) than late eluted proteins<br />

(< 0.24). The reason can be that proteins with high polar residues retained for a shorter space<br />

of time than those with low polar residues. In a previous study, it has been mentioned that<br />

membrane proteins had low polarity below 40% while soluble proteins had higher polarity<br />

89


Chapter 3<br />

approximately 47%. The membrane proteins with the low polarity need detergents to be<br />

separated from respective membranes due to hydrophobic interactions (117). However, in the<br />

same study the polarity has not been correlated with the protein retention behavior in HIC.<br />

The polarity values for the same protein on Grantham’s scale were almost one digit higher<br />

than Zimmerman’s scale. This could be due to the different approaches used by the two<br />

researchers while calculating the polarity indices for individual amino acids. It has been<br />

reported that polar residues are normally located in the outer part of the protein while nonpolar<br />

residues in the inner portion (117). The polarity of a protein has a direct relationship<br />

with the number of polar residues (Asp, Asn, Glu, Lys, Arg and Gln) in the protein sequence.<br />

An inverse relationship was reported between polarity and retention behavior of the proteins<br />

(Figure 25 A-B). This revealed that the proteins with less polarity have a more non-polar<br />

surface area and resulted in stronger binding to the HIC surface. In contrast, proteins with<br />

high polarity had more polar and less non-polar surface area, resulting in less binding to<br />

hydrophobic adsorbents and earlier elution during chromatography. This also ensured that<br />

polarity has an inverse relationship with protein hydrophobicity.<br />

90


Chapter 3<br />

Figure 25 A: Represents the average polarity calculated by Grantham’s scale. The protein<br />

average polarity values of the three base supports were correlated with the respective salt<br />

concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average polarity values in each fraction. The statistical analysis revealed a<br />

significant correlation (r 2 = 0.84; p < 0.0001).<br />

91


Chapter 3<br />

Figure 25 B: Represents the average polarity calculated by Zimmerman’s scale. The protein<br />

average polarity values of the three base supports were correlated with the respective salt<br />

concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average polarity values in each fraction. The statistical analysis revealed a<br />

significant correlation (r 2 = 0.80; p < 0.0001).<br />

The r 2 values of the polarity with the retention of the proteins were 0.84 and 0.80 for the<br />

scales of Grantham and Zimmerman, respectively. An indirect relationship was also observed<br />

between polarity and flexibility of the proteins. The proteins with high polarity were less<br />

flexible and eluted at the beginning of the chromatography. These results confirmed the effect<br />

of average polarity in the chromatographic separation of proteins with the process proteomics<br />

approach. The polarity of the surface amino acids can also be used as an alternative parameter<br />

to average surface hydrophobicity and can be named as average surface polarity (ASP). The<br />

ASP parameter can be calculated for the tabulated proteins in the same way as ASH was<br />

92


Chapter 3<br />

calculated. The ASP parameter can be used for the prediction of protein retention time in HIC.<br />

The average polarity has revealed no differences among the base supports for their<br />

hydrophobicity.<br />

3.2.3.3.3. Average surface hydrophobicity (ASH)<br />

There is no universally accepted method to be considered for the measurement of protein<br />

hydrophobicity; however it can be estimated through several hydrophobicity scales proposed<br />

by different co-workers (25, 92, 95). The hydrophobicity scale proposed by Cowan-Whittaker<br />

(92) was used to calculate ASH. This scale was based on a direct method and was derived<br />

from the amino acid retention time in RP-HPLC. The approach of J.C. Salgado et al (32, 92)<br />

was used for calculating ASH utilizing the bioinformatics tools. This property was<br />

demonstrated by averaging the hydrophobic potentials of the surface amino acid residues<br />

present in the three dimensional structure of a protein. All the amino acids on the surface of a<br />

protein have contributed in the ASH; however the eight non polar hydrophobic amino acids<br />

such as alanine, leucine, isoleucine, tryptophan, methionine, phenylalanine and proline have<br />

the major contributions. The ASH values have been calculated for the tabulated proteins<br />

(Tables 12-14). The statistical analysis revealed a significant direct correlation (r 2 = 0.91, p <<br />

0.002) between ASH and protein retention behavior on the respective adsorbents (Figure 26).<br />

The proteins were classified into three groups depending on the ASH and the corresponding<br />

salt concentrations on which the proteins were eluted. The proteins eluted at high salt<br />

concentrations in the beginning of the chromatography have less ASH values and were<br />

considered as less hydrophobic than the proteins eluted at the middle and / or at the end of the<br />

chromatography. Those proteins which have many hydrophobic residues on their surfaces<br />

were retained longer and eluted at the end of the chromatography. The data obtained for the<br />

proteins eluted at low salt concentrations such as 0.6 M, 0.2 M and 0 M have high importance<br />

for industrial applications, where the use of low salts is the preferred choice in HIC (24). ASH<br />

93


Chapter 3<br />

values have been reported on average 0.403, 0.449 and 0.519 for the proteins eluted at the<br />

beginning, middle and at the end of the chromatography, respectively. These hydrophobicity<br />

values were in agreement to those in the literature (32, 62). This revealed that proteins with<br />

less ASH would retain for a shorter space of time than the protein of high surface<br />

hydrophobicity. The surface hydrophobicity can also be related to the hydrophobic surface<br />

area of a protein and both have linear relations between them and the retention volume (127).<br />

Figure 26: Represents the average surface hydrophobicity calculated by Cowan-Whittaker’s<br />

scale. The protein average surface hydrophobicity values of the three base supports were<br />

correlated with the respective salt concentrations where the fractions were collected. The error<br />

bars represent the standard deviations of the average surface hydrophobicity values in each<br />

fraction. The statistical analysis on average revealed a significant correlation (r 2 = 0.91; p <<br />

0.002).<br />

94


Chapter 3<br />

The correlation coefficients (r 2 ) of ASH with retention of proteins were 0.93, 0.83 and 0.96<br />

for Toyopearl-Phenyl, Sepharose-Phenyl and Source-Phenyl, respectively. The correlation<br />

coefficient reported in the Sepharose-Phenyl case was less than Toyopearl-Phenyl and<br />

Source-Phenyl; the reason could be the heterogeneous distribution of the hydrophobic<br />

hotspots on the base support. This kind of behavior was previously reported with<br />

chromatographic separation of model proteins by Sepharose-Phenyl, where proteins of similar<br />

ASH have exhibited different retention times (63).<br />

In the previous reports, model proteins have been used to correlate ASH with the retention<br />

behavior. This correlation was never examined as a function of the complex cell proteome (26,<br />

27). The correlation of ASH with retention behavior was explored for the first time in this<br />

work with the yeast cell proteome. The experimental data about salt concentrations and<br />

corresponding hydrophobicity ranges at different elution stages can be exploited to predict the<br />

separation of recombinant proteins employing yeast as a host cell proteome. The separation of<br />

recombinant proteins by HIC can also be predicted in other host cell proteomes, once its ASH<br />

is calculated. ASH was the only property which has revealed the trend lines to partially<br />

differentiate among the base supports.<br />

3.2.3.3.4. Other protein properties affecting the chromatographic behavior<br />

The average flexibility was calculated for the tabulated proteins using the scale proposed by<br />

Bhaskaran and Ponnuswamy (Tables 12-14) (98). The average flexibility has revealed a<br />

statistically significant (r 2 = 0.80; p < 0.0001) direct correlation with the protein retention<br />

during chromatography (Figure 27). The average flexibility has been reported before for its<br />

direct relationship with the retention time of model proteins in HIC (62). This could be due to<br />

the presence of cystein residues in a protein forming disulfide bridges. A protein with more<br />

cystein content / and or disulfide bridges will be highly stable and less flexible towards<br />

conformational changes. In contrast, proteins with less cystein residues will be highly flexible<br />

95


Chapter 3<br />

towards conformational changes and when it comes in contact of HIC surface, it will unfold<br />

or expose the internal surface in order to increase hydrophobic interactions and result in<br />

longer retention during chromatography (33, 62, 128).<br />

Figure 27: Represents the average flexibility calculated by Bhaskaran and Ponnuswamy’s<br />

scale. The protein average flexibility values of the three base supports were correlated with the<br />

respective salt concentrations where the fractions were collected. The error bars represent the<br />

standard deviations of the average flexibility values in each fraction. The statistical analysis<br />

revealed a significant correlation (r 2 = 0.80; p < 0.0001).<br />

It was also reported by Tanford that proteins with native conformations were completely<br />

inflexible and conformational entropy was zero. In contrast, the unfolded proteins have high<br />

flexibility towards conformational changes (25). This revealed that proteins with less<br />

flexibility had less conformational entropy and high stability. In this work, average flexibility<br />

was correlated with the retention of proteins derived from a host cell proteome. Some proteins<br />

96


Chapter 3<br />

such as D3-SP, B4-PS and B4-SP were eluted earlier during chromatography than E1-SP, C1-<br />

PS and C1-SP, respectively although they have almost similar ASH. This could be due to the<br />

higher flexibilities of the proteins which allow them to retain longer during chromatography.<br />

The same behavior was also reported with model proteins where ribonuclease S has shown<br />

higher retention on Sepharose-Butyl than other proteins, although they had almost identical<br />

surface hydrophobicity. It was claimed that higher flexibility of ribonulcease S might be<br />

favoring longer retention (63). The highly flexible proteins have revealed more unfolding and<br />

most of the hydrophobic residues were exposed which has given tight binding and in result<br />

most of them eluted at the end of the chromatography. Less flexible proteins were not able to<br />

stay longer due to their compact nature and most of hydrophobic parts were in the inner core,<br />

which resulted in less hydrophobic interactions and elution occurred at the beginning of the<br />

chromatography. Although the flexibility has no decisive role in retention, however it can be<br />

considered as the contributing parameter during HIC. The flexibility has given a correlation<br />

with retention of proteins; however it has not exhibited any differences among the base<br />

supports for their hydrophobic characters.<br />

The average bulkiness was also calculated for the tabulated proteins using Zimmerman’s scale<br />

(Tables 12-14). The average bulkiness of the proteins has revealed no significant relationship<br />

(r 2 = 0.04, p = 0.38) with the chromatographic separation of proteins in HIC (Figure 28). The<br />

bulkiness has been never reported for its contribution in HIC. In the previous report,<br />

Zimmerman had investigated to correlate polarity and bulkiness of cytochrome c protein in<br />

different organisms (96). However, no significant trend was observed between polarity and<br />

bulkiness in his studies. This also confirmed that if bulkiness has no relation with the polarity<br />

of the proteins, then it is supposed to have no relation with protein hydrophobicity as<br />

evidenced here. The bulkiness has also exhibited no differences among the different base<br />

supports.<br />

97


Chapter 3<br />

Figure 28: Represents the average bulkiness calculated by Zimmerman’s scale. The protein<br />

average bulkiness values of the three base supports were correlated with the respective salt<br />

concentrations where the fractions were collected. The error bars represent the standard<br />

deviations of the average bulkiness values in each fraction. The statistical analysis revealed a<br />

non significant correlation (r 2 = 0.04; p = 0.38).<br />

3.2.3.4. Comparative studies of the different base supports based on average protein<br />

properties<br />

The comparison of the base supports has been done for the first time in this work with the<br />

proteome wide approach. The protein properties such as average surface hydrophobicity,<br />

average hydrophobicity, average flexibility and average polarity have shown a correlation<br />

with the retention of proteins during chromatography. However, these properties have a<br />

limited ability to differentiate among the base supports for their hydrophobicity (Table 15).<br />

98


Chapter 3<br />

Table 15: Average protein properties for their correlation with retention time and the ability to<br />

differentiate among the base support for their hydrophobicity.<br />

Protein properties<br />

Differentiation<br />

Correlation<br />

Average hydrophobicity × +<br />

Average polarity ×<br />

_<br />

Average flexibility × +<br />

Average bulkiness × ×<br />

Average surface hydrophobicity Trend +<br />

+ indicates positive correlation, - indicates negative correlation; × indicates no correlation<br />

These results revealed that protein properties have a limited ability to differentiate among the<br />

base supports for their hydrophobicity. At any specific fraction of time, several proteins are<br />

supposed to be eluted during chromatographic separation of a cell proteome. Due to this<br />

reason, assigning a specific retention time to any protein is impractical. This revealed that the<br />

proteome wide approach has a limited ability to differentiate among the base supports.<br />

In two cases (1.6 M and 0 M), the Sepharose-Phenyl has given slightly high hydrophobic<br />

proteins than Toyopearl-Phenyl. This unexpected behavior of Sepharose-Phenyl was also<br />

reported before; where proteins with high ASH were eluted earlier than those with lower ASH.<br />

The reason could be the heterogeneous distribution of hydrophobic hotspots on the<br />

Sepharose-Phenyl, due to which an unusual behavior was reported in the retention mechanism<br />

(63). Another observation was that base supports revealed less differences at high salt<br />

concentrations than at low salt concentrations. In other words, the differences among the base<br />

supports were less at high hydrophobic conditions than at low hydrophobic conditions. This<br />

kind of behavior was also reported in our group with model proteins, where the differences<br />

99


Chapter 3<br />

among the adsorbents were higher at low salt concentrations than at the high salt<br />

concentrations (121). The reason for this could be that at the extreme hydrophobic conditions,<br />

the hydrophobic potential of the mobile phase has covered the potential of the base supports<br />

to demonstrate according to their hydrophobicities. At low salt concentrations, the mobile<br />

phase has low hydrophobicity and the adsorption was more dependent on the hydrophobicity<br />

of the base supports.<br />

3.2.3.5. Application note<br />

This work can help those interested in the purification of proteins listed in Tables 12-14. It<br />

will simplify the work of researchers and they will not have to scout for chromatographic<br />

experiments to figure out the elution position of any tabulated protein. The protein can be<br />

purified by isocratic chromatography, without going through trial and error chromatographies.<br />

Some of the important enzymes are Pyruvate kinase (eluted at 0.6 M on Sepharose-Phenyl),<br />

Phosphoglycerate mutase 2 (eluted at 0.8 M on Toyopearl-Phenyl), Polymerase A (0 M on<br />

Source-Pheny), Methyl transferase (0.2 M on Sepharose-Phenyl), Thiosulfate sulphur<br />

transferase (0.2 M on Source-Phenyl), Enolase 2 (1.6 M on Toyopearl-Phenyl), D-serine<br />

dehydratase (0.6 M on Toyopearl-Phenyl) and Alcohol dehydrogenase (0.2 M on Sepharose-<br />

Phenyl).<br />

100


Chapter 3<br />

3.2.3. Partial conclusions<br />

• The influence of the base supports has been investigated for the first time with the<br />

proteome wide approach utilizing different protein properties. The results revealed<br />

that protein properties have limited ability to differentiate among the base supports.<br />

• The base supports revealed less difference at high salt concentrations than at low salt<br />

concentrations. In other words, the differences among the base supports were less at<br />

high hydrophobic conditions than at low hydrophobic conditions. The reason could be<br />

the effects of mobile phase hydrophobicity.<br />

• Several protein properties such as average hydrophobicity, average flexibility and<br />

average surface hydrophobicity have revealed direct relationships with the retention of<br />

proteins. Average polarity has given an inverse relationship to the retention of proteins<br />

during HIC.<br />

• Among all the properties, ASH was the only property which has a decisive role in the<br />

retention behavior of proteins during chromatography.<br />

• The data collected from the three base supports in this process proteomics approach<br />

can be used to design the purification models by the in silico approach.<br />

101


Chapter 3<br />

3.3. Protein distributions on 2-D gels: the influence of molecular<br />

weight and isoelectric point coordinates<br />

3.3.1. Summary<br />

The separation behavior of the yeast cell proteome by hydrophobic interaction<br />

chromatography (HIC) was examined utilizing two dimensional polyacrylamide gel<br />

electrophoresis (2-D PAGE). The yeast cell proteome was fractionated on different<br />

hydrophobic adsorbents and chromatographic fractions were collected on specific salt<br />

concentrations. The chromatographic fractions were further analyzed by 2-D PAGE to<br />

investigate the contaminant profile in each chromatographic fraction. This chapter will give<br />

an additional understanding to the influence of ligands and base supports on the<br />

chromatographic separation of the yeast cell proteome. As expected, Toyopearl-Hexyl and<br />

Toyopearl-Butyl have shown increased protein retention in comparison with Toyopearl-Ether.<br />

Moreover, an influence of protein size and isoelectric point (pI) was also revealed as a<br />

function of the adsorbent type. Larger and / or neutral proteins showed increased while acidic<br />

(pI < 6) and basic (pI > 8) proteins depicted decreased or increased retention, depending on<br />

the adsorbent type.<br />

In addition to the separation behavior of yeast cell proteome during chromatography, an<br />

overview of the yeast cell proteome was gained. A high amount and a higher number of the<br />

proteins on 2-D gels were reported at the beginning of the chromatography than at the end.<br />

This confirmed the less hydrophobic nature of the yeast cell proteome. A higher number of<br />

smaller and basic proteins were observed on 2-D gels in comparison to larger and acidic<br />

proteins. This further showed the nature of the yeast cell proteome.<br />

102


Chapter 3<br />

3.3.2. Results and Discussion<br />

3.3.2.1. Unfolding of proteins during HIC<br />

The 2-D gels were performed for all the chromatographic fractions collected through six<br />

different hydrophobic adsorbents and stained with colloidal blue method in order to get the<br />

precise number of spots available in each fraction. The 2-D gels have been presented in<br />

Figures 12-14 (Section 3.1) and Figures 21-23 (Section 3.2). Some denatured spots were<br />

observed for the fractions collected at high salt concentrations on Source-Phenyl and<br />

Toyopearl-Hexyl representing the conformational changes occurred during chromatography.<br />

However, Toyopearl-Butyl and Sepharose-Phenyl has revealed less conformational changes<br />

and solid spots were observed on the 2-D gels. The conformational changes at high salt<br />

concentrations during chromatography were also reported before with model proteins (122,<br />

129-131). The ratio of the conformational changes of proteins was high on the gels collected<br />

at the high salt concentrations and extreme hydrophobic adsorbents. This revealed that<br />

conformational changes during chromatography have dependence on the salt concentration<br />

and the adsorbent type. This behavior was also reported before that by increasing the<br />

ammonium sulfate concentration and the chain length of ligands, the unfolding of protein<br />

increases on the HIC surface. This is the reason that unnecessary high salt concentration and<br />

high chain lengths of the ligands should not be preferred (129). The conformational changes<br />

observed on 2-D gels further reinforced the correlations observed in Section 3.1 and Section<br />

3.2, between flexibility and the corresponding retention volume. Some protein loss was also<br />

observed while comparing the amount of sample applied during chromatography and proteins<br />

obtained in the fractions <strong>after</strong> chromatography. Small amount of proteins (almost 5 mg) was<br />

lost elsewhere on the hydrophobic surface. The reason could be the unfolding of the proteins<br />

which resulted in less recovery in the chromatography fractions. These results confirmed the<br />

protein unfolding during HIC with the proteome wide approach.<br />

103


Chapter 3<br />

3.3.2.2. Protein distributions on 2-D gels<br />

Proteins derived from S. cerevisiae cell proteome have shown less binding affinity on the<br />

adsorbents based on hydrophobic interactions. A maximum number and a high amount (mg)<br />

of proteins were eluted at the beginning of the chromatography than at the end (Figure 29 A-<br />

B). These results confirmed the less hydrophobic nature of S. cerevisiae cell proteome. The<br />

hydrophobic proteins in the S. cerevisiae cell proteome were reported to be 16% of the total<br />

cell proteome. This was very less amount than reported for other cell proteomes (104). The<br />

less hydrophobic nature of the yeast cell proteome suggested high hydrophobic conditions for<br />

the efficient separation of the cell proteome during HIC. The analysis of 2-D gels has also<br />

revealed the variation in the binding affinity of the different ligands for the same proteome. In<br />

case of Toyopearl-Ether, a higher number of protein spots were eluted at the beginning of the<br />

chromatography than at the end. The reason could be the less hydrophobic nature of the<br />

Toyopearl-Ether. In contrast, Toyopearl-Hexyl has shown higher numbers of protein spots at<br />

the end of the chromatography than at the beginning because of its high hydrophobic nature.<br />

In the similar fashion, the 2-D gels analysis has also revealed the variation in the binding<br />

affinity of the base supports for the same proteome. In case of Toyopearl-Phenyl, a higher<br />

number of proteins were eluted in the beginning of the chromatography than at the end and<br />

the reason could be the less hydrophobic nature of the base support. Sepharose-Phenyl and<br />

Source-Phenyl have shown higher numbers of spots at the end of the chromatography. This<br />

revealed the high binding affinity of the two adsorbents for the yeast cell proteome.<br />

When the adsorption affinity of polymethacrylate based media was compared with phenyl<br />

ligated media, the adsorption affinity was higher in earlier than later one. Toyopearl-Phenyl<br />

with phenyl ligand has not given higher adsorption of proteins when compared to Toyopearl-<br />

Butyl and Toyopearl-Hexyl. This confirmed that aromatic interactions have no effect on the<br />

adsorption affinity of the adsorbents (122, 132).<br />

104


Chapter 3<br />

Figure 29 A: Represent the total number of spots and amount of proteins (mg) in each fraction<br />

as a function of the different base supports. The number of protein spots and amount of<br />

proteins (mg) revealed a significant correlation with the respective salt concentrations where<br />

the fractions were collected.<br />

105


Chapter 3<br />

Figure 29 B: Represent the total number of spots and amount of proteins (mg) in each fraction<br />

as a function of the different ligands. The number of protein spots and amount of proteins (mg)<br />

revealed a significant correlation with the respective salt concentrations where the fractions<br />

were collected.<br />

3.3.2.3. The effect of protein molecular mass on the chromatographic behavior<br />

Other component in evaluating the experimental data obtained <strong>after</strong> 2-D gel electrophoresis is<br />

the potential relationship between protein molecular weight and elution behavior, as a<br />

function of elution position and adsorbent type. In a previous report, molecular mass of the<br />

protein has not shown any correlation with protein retention in ion exchange chromatography<br />

(19). However, the effect of molecular mass was reported in HIC using the retention time of<br />

the model proteins (33). A few model proteins were used for the analysis which may not give<br />

a complete picture of all the contributing parameters involved in protein retention. In this<br />

106


Chapter 3<br />

work, the effects of larger proteins (> 50 KDa) on retention have been examined as a function<br />

of the adsorbent type. The number of larger proteins in each fraction was divided by the total<br />

number of proteins to normalize the data. The normalized values of the larger proteins from<br />

the three ligands was presented in Figure 30, where the larger proteins were found to be<br />

higher at low salt concentrations than at high salt concentrations. In other words, the larger<br />

proteins were higher in the latter fractions in comparison to the earlier ones. This revealed that<br />

larger proteins retained longer on the adsorbents. In case of larger proteins, the retention was<br />

higher for Toyopearl-Butyl and Toyopearl-Hexyl in comparison to the other adsorbents. It<br />

was reported before that if the pore space of the adsorbent is smaller than the dimensions of<br />

the magnitude as the protein has, then it will result in less retention of protein (33). However,<br />

this was not applicable on Toyopearl-Ether and Toyopearl-Phenyl adsorbents where larger<br />

proteins has shown less retention, although they have high pore size (100 nm). The reason<br />

could be the less hydrophobic nature of the two adsorbents which retained larger proteins for<br />

a shorter time during chromatography. In a similar way, Source-Phenyl has given less<br />

retention to larger proteins, although it has a high hydrophobic nature, but the pore size is<br />

smaller (34 nm). These results declared that the pore size and hydrophobicity of the<br />

adsorbents could be the combine contributing parameters affecting the retention of larger<br />

proteins.<br />

In case of smaller proteins it was reported once that it has no effect on retention (33).<br />

However, in another report it was reported that smaller proteins exhibited a decrease in<br />

retention due to the less contact surface area with adsorbents (68). This was confirmed in this<br />

work when smaller proteins have exhibited an inverse relationship to the corresponding<br />

retention behavior and most of the smaller proteins were eluted at the beginning of the<br />

chromatography. The reason could be that smaller proteins have less non-polar surface area to<br />

interact with the adsorbents.<br />

107


Chapter 3<br />

When the number of smaller and larger proteins was compared, smaller proteins were three<br />

times higher than larger proteins on the 2-D gels. This revealed that S. cerevisiae cell<br />

proteome have a higher proportion of smaller proteins than larger proteins.<br />

Figure 30: Represents the normalized values of the larger proteins (Mr > 50) in each fraction as<br />

a function of the different ligands. The larger proteins revealed a significant correlation with<br />

the respective salt concentrations where the fractions were collected.<br />

3.3.2.3. The influence of protein isoelectric point on the chromatographic behavior<br />

The effect of mobile phase pH and isoelectric point of the proteins are the parameters least<br />

studied in HIC. At the start of HIC method, it was reported that there is no obvious<br />

relationship between retention and isoelectric point of individual proteins (68). However, in<br />

the follow up studies some clear pH effects were observed, where basic proteins has shown<br />

longer retention when the mobile phase pH was close to their pI and acidic proteins exhibited<br />

less retention with basic pH values. In other words, the protein retention was increased at<br />

108


Chapter 3<br />

acidic pH and decreased at basic pH values of the mobile phase (43). To study the sole effect<br />

of the pI, a neutral pH was maintained during chromatographic experiments. The neutral pH<br />

is also a preferred choice in the downstream processing of proteins. The chromatographic<br />

fractions were analyzed on 2-D gels and the normalized values of the acidic and basic proteins<br />

were calculated. The normalized values of the basic proteins in each gel (fraction) were<br />

presented versus corresponding elution position for different ligands (Figure 31). A higher<br />

number of basic proteins were found at low salt concentrations at the end of the<br />

chromatography which revealed the longer retention of basic proteins during chromatography.<br />

The Sepharose-Phenyl and Toyopearl-Butyl have given longer retention to the basic proteins<br />

in comparison to the other adsorbents. This revealed the partial negative charges on these<br />

adsorbents. The negative charges on adsorbents may attract the positive charges on basic<br />

proteins and resulted in ionic interactions in addition to HIC. The combination of the ionic<br />

and hydrophobic interactions resulted in longer retention of the basic proteins during<br />

chromatography. The partial negative charges on hydrophobic adsorbents were also reported<br />

in literature, although the ration of negative charges varied among adsorbents (133). The<br />

uncharged adsorbents were reported to be preferred for the true hydrophobic interactions,<br />

otherwise at low salt concentrations; the interaction will be more electrostatic than<br />

hydrophobic, which can affect the separation based on hydrophobic interactions (43).<br />

However, in other reports the presence of charges on the adsorbents was favored in order to<br />

maintain the native structures of the proteins (21, 102).<br />

The neutral proteins pI (6-8) and acidic proteins (pI < 6) have been eluted in almost equal<br />

distribution during the chromatographic separations. The reason could be the neutral pH of<br />

the mobile phase, which allowed the neutral proteins to elute on the basis of hydrophobicity.<br />

The analysis of 2-D gels revealed that the gels at the end have mostly neutral proteins, while<br />

acidic and basic proteins are rarely available. The proteins with pI around 6, 7 and 8 have<br />

been highly retained at neutral pH and the reason could be that the charge was near to zero<br />

109


Chapter 3<br />

and hydrophobic interactions were at the maximum extent between proteins and adsorbents<br />

(44). These results confirmed the effects of isoelectric point values of the proteins on the<br />

chromatographic separation as a function of adsorbent type. A high number of basic proteins<br />

have been reported on 2-D gels in comparison to acidic proteins. This revealed the high<br />

abundance of basic proteins in S. cerevisiae cell proteome.<br />

Figure 31: Represents the normalized values of the basic proteins (pI > 8) in each fraction as a<br />

function of different ligands. The basic proteins revealed a significant correlation with the<br />

respective salt concentrations where the fractions were collected.<br />

110


Chapter 3<br />

3.3.3. Partial conclusions<br />

• A high amount and a higher number of protein spots were reported at the beginning of<br />

the chromatography than at the end. This confirmed the less hydrophobic nature of the<br />

S. cerevisiae cell proteome.<br />

• The denaturation of proteins was confirmed in HIC with the proteome wide approach.<br />

• Aromatic interactions were confirmed to have no effect on the retention behavior of<br />

proteins during HIC.<br />

• The effect of molecular weight of the proteins was studied in relation to the adsorbent<br />

type. The pore size and hydrophobicity of the adsorbents were reported as the main<br />

parameters responsible for the retention of the larger proteins.<br />

• The effect of the isoelectric point of proteins was studied in relation to adsorbent type.<br />

The neutral and acidic proteins have given equal distributions over 2-D gels. The<br />

partial negative charges were confirmed on the hydrophobic adsorbents which has<br />

given rise to ionic interactions in addition to hydrophobic interactions and resulted in<br />

longer retention of basic proteins during chromatography.<br />

• A higher number of acidic and smaller proteins were reported in the cell proteome in<br />

comparison to basic and larger proteins, which can further reflect the nature of S.<br />

cerevisiae cell proteome.<br />

111


Chapter 4<br />

CONCLUSIONS


Chapter 4<br />

4. Conclusions and Remarks<br />

4.1. General conclusions and remarks<br />

This thesis explored for the first time the comparative analysis of hydrophobic ligands and<br />

base supports in hydrophobic interaction chromatography (HIC) as a function of complex cell<br />

proteome. Several protein properties were examined for their correlation with the protein<br />

retention behavior in HIC. The protein properties were also explored to determine their<br />

potential to differentiate among the ligands and base supports. The previous studies of HIC<br />

were based on the retention time of model proteins. Hence, a process proteomics approach<br />

was applied to explore the hidden underlying parameters in HIC. Although very laborious and<br />

experimentally demanding, the present approach has led to conclusions which are valid for a<br />

complex cell proteome and which refers to two families of hydrophobic adsorbents. This<br />

work has produced a first draft of the experimental data which can be used to predict the<br />

elution position of any target protein utilizing an in silico approach.<br />

4.2. Influence of different ligands<br />

• The hydrophobicity of the ligands was comparatively investigated in the previous<br />

reports exploiting the retention time of model proteins. However, the hydrophobicity<br />

of the ligands has been investigated for the first time with the yeast cell proteome<br />

utilizing several protein properties. The protein properties have not fully differentiated<br />

among the ligands. This confirmed that protein properties have limited ability to<br />

differentiate among the ligands for their hydrophobicity.<br />

4.3. Influence of different base supports<br />

• The influence of the base supports has been investigated for the first time with the<br />

proteome wide approach utilizing different protein properties. The results revealed that<br />

protein properties have limited ability to differentiate among the base supports.<br />

113


Chapter 4<br />

• The base supports revealed less difference at high salt concentrations than at low salt<br />

concentrations. In other words, the differences among the base supports were less at<br />

high hydrophobic conditions than at low hydrophobic conditions. The reason could be<br />

the effects of mobile phase hydrophobic conditions.<br />

4.4. Effect of protein properties on separation behavior during chromatography<br />

• Several protein properties were investigated for the first time with the proteome wide<br />

approach for their effects on the protein retention behavior during HIC.<br />

• The average surface hydrophobicity (ASH), average hydrophobicity (AH) and average<br />

flexibility (AF) have revealed direct relationships with the retention behavior of the<br />

proteins in HIC.<br />

• The average bulkiness of the proteins did not show any relationship with the retention<br />

behavior during chromatography.<br />

• The polarity of the proteins exhibited an inverse relationship with the retention of the<br />

proteins. The higher the polarity of the protein, the less time it will stay on the<br />

adsorbent during chromatography and vice versa.<br />

• An indirect relationship was observed between polarity and hydrophobicity; proteins<br />

with high polarity were less hydrophobic in nature and vice versa.<br />

• An indirect relationship was also observed between polarity and flexibility; proteins<br />

with high polarity were less flexible in nature and vice versa.<br />

• Among all the properties, ASH was the only property which has a decisive role in the<br />

retention behavior of the proteins during HIC.<br />

• The salt and ASH ranges were defined for the first time during the chromatographic<br />

fractionation of the yeast cell proteome. The information gathered about ASH and the<br />

retention position during chromatography can be exploited for the chromatographic<br />

separation of recombinant proteins from host cell proteomes.<br />

114


Chapter 4<br />

• The experimental data obtained with this proteome wide approach can be used to<br />

design the purification models utilizing the in silico approach.<br />

4.5. 2-D gels analysis<br />

• A high amount and a higher number of protein spots were reported at the beginning of<br />

the chromatography than at the end. This result confirmed the less hydrophobic nature<br />

of the S. cerevisiae cell proteome.<br />

• A higher number of smaller and basic proteins were reported than larger and acidic<br />

proteins which can further reflect the nature of the S. cerevisiae cell proteome.<br />

• The denaturation of proteins in HIC was confirmed with the proteome wide approach.<br />

• A large number of proteins were retained on Toyopearl-Butyl than Toyopearl-Phenyl,<br />

this confirmed that aromatic interactions have no effect on the retention behavior<br />

during HIC.<br />

• The effect of molecular weight of the proteins was studied in relation to the adsorbent<br />

type. The pore size and hydrophobicity of the adsorbents were concluded to be the<br />

main parameters responsible for the retention of larger proteins in HIC.<br />

• The effect of isoelectric point of the proteins was studied in relation to the adsorbent<br />

type. The neutral and acidic proteins were equally distributed over the 2-D gels. Basic<br />

proteins retained longer due to the presence of partial negative charges on the<br />

adsorbents which has given rise to ionic interactions in addition to hydrophobic<br />

interactions and resulted in longer retention of basic proteins.<br />

4.5. Additional conclusions<br />

• In total, 173 proteins were identified and the in-gel digestion protocol was optimized<br />

specifically for the proteins fractionated by HIC. The identified proteins can provide<br />

115


Chapter 4<br />

additional understanding of the protein contaminants present in the yeast cell proteome<br />

and is a valuable addition to the proteomics technology.<br />

• Several important enzymes have been identified which were eluted at specific salt<br />

concentrations. These enzymes can be purified by a single isocratic chromatography<br />

without going through trial and error chromatographies to figure out the elution<br />

position of any enzyme.<br />

4.6. Recommendations for future work<br />

• The average protein properties derived from the primary structure need to be explored<br />

in reverse phase chromatography with the process proteomics approach.<br />

• The average polarity parameter can be calculated from the surface residues of the three<br />

dimensional structure of a protein. This average surface polarity can be used as an<br />

alternative parameter to ASH.<br />

• The adsorbent such as Toyopearl-Ether revealed an interesting chromatographic<br />

behavior and can be used in future for the purification of the recombinant proteins of<br />

highly hydrophobic nature.<br />

• This approach can be extended to other hydrophobic adsorbents and expression<br />

systems such as E.coli, mammalian cells and insect cells for the development of<br />

downstream processing.<br />

• This work involved the analysis of a real expression host instead of using model<br />

proteins and the data obtained can be used to design a purification model for the<br />

recombinant proteins utilizing an in silico approach.<br />

116


References<br />

5. References<br />

1. Singh, B.D. Biotechnology: expanding horizons. Kalyani: 2008.<br />

2. Ratledge, C. and Kristiansen, B. Basic biotechnology. Cambridge <strong>University</strong> Press: 2006.<br />

3. Pavlou, A.K. and Reichert, J.M. Recombinant protein therapeutics-success rates, market<br />

trends and values to 2010. Nature Biotechnology 2004, 22, 1513-1519.<br />

4. Harrison, R.G. Bioseparations science and engineering: 2003.<br />

5. Headon, D.R. and Walsh, G. The industrial production of enzymes. Biotechnology<br />

Advances 1994, 12, 635-646.<br />

6. Wiseman, A. Designer enzyme and cell applications in industry and in environmental<br />

monitoring. Journal of Chemical Technology and Biotechnology 1993, 56, 3-13.<br />

7. Labrou, N. and Clonis, Y.D. The affinity technology in downstream processing. Journal of<br />

Biotechnology 1994, 36, 95-119.<br />

8. Wheelwright, S.M. Protein purification: design and scale up of downstream processing:<br />

1991.<br />

9. Lienqueo, M.E. and Asenjo, J.A. Use of expert systems for the synthesis of downstream<br />

protein processes. Computers & Chemical Engineering 2000, 24, 2339-2350.<br />

10. Hooker, B.S. and Lee, J.M. Application of a new structured model to tobacco cell cultures.<br />

Biotechnology and bioengineering 1992, 39, 765-774.<br />

11. Asenjo, J.A.; Herrera, L.; Byrne, B. Development of an expert system for selection and<br />

synthesis of protein purification processes. Journal of Biotechnology 1989, 11, 275-298.<br />

12. Eriksson, H.; Sandahl, K.; Brewer, J.; Osterlund, B. Reactive planning for<br />

chromatography. Chemometrics and Intelligent Laboratory Systems 1991, 13, 185-194.<br />

117


References<br />

13. Lienqueo, M.E.; Leser, E.W.; Asenjo, J.A. An Expert System for the selection and<br />

synthesis of multistep protein separation processes. Computers & Chemical Engineering 1996,<br />

20, Supplement 1, S189-S194.<br />

14. Buydens, L.M.C.; van Leeuwen, J.A.; Mulholland, M.; Vandeginste, B.G.M.; Kateman, G.<br />

An expert system for the validation of high-performance liquid chromatographic methods.<br />

TrAC Trends in Analytical Chemistry 1990, 9, 58-62.<br />

15. Scott M., W. The design of downstream processes for large-scale protein purification.<br />

Journal of Biotechnology 1989, 11, 89-102.<br />

16. Watanabe, E.; Tsoka, S.; Asenjo, J.A. Selection of chromatographic protein purification<br />

operations based on physicochemical properties. Annals of the New York Academy of Sciences<br />

1994, 721, 348-364.<br />

17. Lowe, C.R.; Lowe, A.R.; Gupta, G. New developments in affinity chromatography with<br />

potential application in the production of biopharmaceuticals. Journal of Biochemical and<br />

Biophysical Methods 2001, 49, 561-574.<br />

18. Novais, J.L.; Titchener-Hooker, N.; Hoare, M. Economic comparison between<br />

conventional and disposables-based technology for the production of biopharmaceuticals.<br />

Biotechnology and bioengineering 2001, 75, 143-153.<br />

19. Cabrera, R. and Fernandez-Lahore, M. Global screening of protein chromatographic<br />

behavior on ion exchangers from a complex cell proteome: Towards in silico downstream<br />

processing of bioproducts. Journal of Chromatography A 2007, 1161, 41-50.<br />

20. GE healthcare Hydrophobic interactions and reverse phase chromatography Principles and<br />

methods hand book 11-0012-69AA.<br />

21. Vailaya, A. and Horvath, C. Retention thermodynamics in hydrophobic interaction<br />

chromatography. Industrial & Engineering Chemistry Research 1996, 35, 2964-2981.<br />

22. Ceccaroli, P.; Cardoni, P.; Buffalini, M.; De Bellis, R.; Piccoli, G.; Stocchi, V. Separation<br />

of hexokinase activity using different hydrophobic interaction supports. Journal of<br />

Chromatography B: Biomedical Sciences and Applications 1997, 702, 41-48.<br />

118


References<br />

23. Lienqueo, M.E. and Mahn, A. Predicting protein retention time in hydrophobic interaction<br />

chromatography. Chemical Engineering & Technology 2005, 28, 1326-1334.<br />

24. Kato, Y.; Nakamura, K.; Kitamura, T.; Moriyama, H.; Hasegawa, M.; Sasaki, H.<br />

Separation of proteins by hydrophobic interaction chromatography at low salt concentration.<br />

Journal of Chromatography A 2002, 971, 143-149.<br />

25. Tanford, C. Contribution of hydrophobic interactions to the stability of the globular<br />

confirmation of proteins. Journal of the American Chemical Society 1962, 84, 4240-4247.<br />

26. Mahn, A. and Asenjo, J.A. Prediction of protein retention in hydrophobic interaction<br />

chromatography. Biotechnology Advances 2005, 23, 359-368.<br />

27. Lienqueo, M.E.; Mahn, A.; Salgado, J.C.; Asenjo, J.A. Current insights on protein<br />

behaviour in hydrophobic interaction chromatography. Journal of Chromatography B 2007,<br />

849, 53-68.<br />

28. Lin, F.; Chen, W.; Ruaan, R.; Huang, H. Microcalorimetric studies of interactions<br />

between proteins and hydrophobic ligands in hydrophobic interaction chromatography: effects<br />

of ligand chain length, density and the amount of bound protein. Journal of Chromatography<br />

A 2000, 872, 37-47.<br />

29. Lefort, S. and Ferrara, P. Hydrophobic adsorbants for the isolation and purification of<br />

biosynthetic human growth hormone from crude fermentation mixtures. Journal of<br />

Chromatography A 1986, 361, 209-216.<br />

30. Gueorguieva, L.; Palani, S.; Rinas, U.; Jayaraman, G.; Seidel-Morgenstern, A.<br />

Recombinant protein purification using gradient assisted simulated moving bed hydrophobic<br />

interaction chromatography--Part II: Process design and experimental validation. Journal of<br />

Chromatography A 2011.<br />

31. Mahn, A.; Lienqueo, M.E.; Salgado, J.C. Methods of calculating protein hydrophobicity<br />

and their application in developing correlations to predict hydrophobic interaction<br />

chromatography retention. Journal of Chromatography A 2009, 1216, 1838-1844.<br />

119


References<br />

32. Salgado, J.C.; Rapaport, I.; Asenjo, J.A. Is it possible to predict the average surface<br />

hydrophobicity of a protein using only its amino acid composition? Journal of<br />

Chromatography A 2005, 1075, 133-143.<br />

33. To, B.C.S. and Lenhoff, A.M. Hydrophobic interaction chromatography of proteins: I.<br />

The effects of protein and adsorbent properties on retention and recovery. Journal of<br />

Chromatography A 2007, 1141, 191-205.<br />

34. Porath, J.; Sundberg, L.,; Fornstedt, N.; Olson, I. Salting-out in amphiphilic gels as a new<br />

approach to hydrophobia adsorption. Nature 1973, 245, 465-466.<br />

35. Hofstee, B.H.J. and Frank Otillio, N. Modifying factors in hydrophobic protein binding by<br />

substituted agaroses. Journal of Chromatography A 1978, 161, 153-163.<br />

36. Stellan, H. Some general aspects of hydrophobic interaction chromatography. Journal of<br />

Chromatography A 1973, 87, 325-331.<br />

37. Melander, W.R.; Corradini, D.; Horváth, C. Salt-mediated retention of proteins in<br />

hydrophobic-interaction chromatography : Application of solvophobic theory. Journal of<br />

Chromatography A 1984, 317, 67-85.<br />

38. Xia, F.; Nagrath, D.; Garde, S.; Cramer, S.M. Evaluation of selectivity changes in HIC<br />

systems using a preferential interaction based analysis. Biotechnology and bioengineering<br />

2004, 87, 354-363.<br />

39. Melander, W. and Horváth, C. Salt effects on hydrophobic interactions in precipitation<br />

and chromatography of proteins: An interpretation of the lyotropic series. Archives of<br />

Biochemistry and Biophysics 1977, 183, 200-215.<br />

40. Porath, J. Salt-promoted adsorption: Recent developments. Journal of Chromatography B:<br />

Biomedical Sciences and Applications 1986, 376, 331-341.<br />

41. Rippel, G. and Szepesy, L. Hydrophobic interaction chromatography of proteins on an<br />

Alkyl-Superose column. Journal of Chromatography A 1994, 664, 27-32.<br />

120


References<br />

42. Halperin, G.; Breitenbach, M.; Tauber-Finkelstein, M.; Shaltiel, S. Hydrophobic<br />

chromatography on homologous series of alkylagaroses : A comparison of charged and<br />

electrically neutral column materials. Journal of Chromatography A 1981, 215, 211-228.<br />

43. Hjertén, S.; Yao, K.; Eriksson, K.; Johansson, B. Gradient and isocratic high-performance<br />

hydrophobic interaction chromatography of proteins on agarose columns. Journal of<br />

Chromatography A 1986, 359, 99-109.<br />

44. Xia, F.; Nagrath, D.; Cramer, S.M. Effect of pH changes on water release values in<br />

hydrophobic interaction chromatographic systems. Journal of Chromatography A 2005, 1079,<br />

229-235.<br />

45. Sanz, Y.; Mulholland, F.; Toldra, F. Purification and Characterization of a Tripeptidase<br />

from Lactobacillus sake. Journal of Agricultural and Food Chemistry 1998, 46, 349-353.<br />

46. Wu, S.; Figueroa, A.; Karger, B.L. Protein conformational effect in hydrophobic<br />

interaction chromatography : Retention characterization and the role of mobile phase<br />

additives and stationary phase hydrophobicity. Journal of Chromatography A 1986, 371, 3-27.<br />

47. Geng, X.; Guo, L.; Chang, J. Study of the retention mechanism of proteins in hydrophobic<br />

interaction chromatography. Journal of Chromatography A 1990, 507, 1-23.<br />

48. Heinitz, M.L.; Kennedy, L.; Kopaciewicz, W.; Regnier, F.E. Chromatography of proteins<br />

on hydrophobic interaction and ion-exchange chromatographic matrices: mobile phase<br />

contributions to selectivity. Journal of Chromatography A 1988, 443, 173-182.<br />

49. Hjertén, S.; Rosengren, J.; Pahlman, S. Hydrophobic interaction chromatography : The<br />

synthesis and the use of some alkyl and aryl derivatives of agarose. Journal of<br />

Chromatography A 1974, 101, 281-288.<br />

50. Parsegian, V.A. and Ninham, B.W. Temperature-Dependent van der Waals Forces.<br />

Journal of Biochemical and Biophysical Methods 1970, 10, 664-674.<br />

51. Queiroz, J.A.; Tomaz, C.T.; Cabral, J.M.S. Hydrophobic interaction chromatography of<br />

proteins. Journal of Biotechnology 2001, 87, 143-159.<br />

121


References<br />

52. B.H.J., H. Accessible hydrophobic groups of native proteins. Biochemical and biophysical<br />

research communications 1975, 63, 618-624.<br />

53. B.H.J., H. Hydrophobic affinity chromatography of proteins. Analatyical Biochemistry<br />

1973, 52, 430-448.<br />

54. Shaltiel, S. and Er-El, Z. Hydrophobic Chromatography: use for purification of glycogen<br />

synthetase. Proceedings of the National Academy of Sciences 1973, 70, 778-781.<br />

55. Hofstee, B.H.J. and Otillio, N.F. Non-ionic adsorption chromatography of proteins.<br />

Journal of Chromatography A 1978, 159, 57-69.<br />

56. Hubert, P.; Mathis, R.; Dellacherie, E. Polymer ligands for mild hydrophobic interaction<br />

chromatography —principles, achievements and future trends. Journal of Chromatography A<br />

1991, 539, 297-306.<br />

57. Jennissen, H.P. and Heilmeyer, L.M.G. General aspects of hydrophobic chromatography.<br />

Adsorption and elution characteristics of some skeletal muscle enzymes. Biochemistry 1975,<br />

14, 754-760.<br />

58. Rosengren, J.; Pahlman, S.; Glad, M.; Hjertén, S. Hydrophobic interaction<br />

chromatography on noncharged sepharose® derivatives: Binding of a model protein, related<br />

to ionic strength, hydrophobicity of the substituent, and degree of substitution (determined by<br />

NMR). Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology<br />

1975, 412, 51-61.<br />

59. Porath, J.; Janson, J.; Ls, T. Agar derivatives for chromatography, electrophoresis and gelbound<br />

enzymes: I. Desulphated and reduced cross-linked agar and agarose in spherical bead<br />

form. Journal of Chromatography A 1971, 60, 167-177.<br />

60. Maisano, F.; Belew, M.; Porath, J. Synthesis of new hydrophobic adsorbents based on<br />

homologous series of uncharged alkyl sulphide agarose derivatives. Journal of<br />

Chromatography A 1985, 321, 305-317.<br />

61. Machold, C.; Deinhofer, K.; Hahn, R.; Jungbauer, A. Hydrophobic interaction<br />

chromatography of proteins: I. Comparison of selectivity. Journal of Chromatography A 2002,<br />

972, 3-19.<br />

122


References<br />

62. Ladiwala, A.; Xia, F.; Luo, Q.; Breneman, C.M.; Cramer, S.M. Investigation of protein<br />

retention and selectivity in HIC systems using quantitative structure retention relationship<br />

models. Biotechnology and bioengineering 2006, 93, 836-850.<br />

63. Mahn, A.; Lienqueo, M.E.; Asenjo, J.A. Effect of surface hydrophobicity distribution on<br />

retention of ribonucleases in hydrophobic interaction chromatography. Journal of<br />

Chromatography A 2004, 1043, 47-55.<br />

64. Pasdar, A.; Nooshin, C.L.; Eunice C. Y. Comparison of protein surface hydrophobicity<br />

measured at various pH values using three different fluorescent probes. Journal of<br />

Agricultural and Food Chemistry 2000, 48, 328-334.<br />

65. Mahn, A.; Lienqueo, M.E.; Asenjo, J.A. Optimal operation conditions for protein<br />

separation in hydrophobic interaction chromatography. Journal of Chromatography B 2007,<br />

849, 236-242.<br />

66. Charles, C, B. On the average hydrophobicity of proteins and the relation between it and<br />

protein structure. Journal of theoretical biology 1967, 16, 187-211.<br />

67. Berggren, K.; Wolf, A.; Asenjo, J.A.; Andrews, B.A.; Tjerneld, F. The surface exposed<br />

amino acid residues of monomeric proteins determine the partitioning in aqueous two-phase<br />

systems. Biochemical at Biophyica Acta-Protein Structure and Molelcular Enzymymology<br />

2002, 1596, 253-268.<br />

68. Fausnaugh, J.L.; Kennedy, L.A.; Regnier, F.E. Comparison of hydrophobic-interaction<br />

and reversed-phase chromatography of proteins. Journal of Chromatography A 1984, 317,<br />

141-155.<br />

69. McNay, J.L.M. and Fernandez, E.J. Protein unfolding during reversed-phase<br />

chromatography: I. Effect of surface properties and duration of adsorption. Biotechnology and<br />

bioengineering 2001, 76, 224-232.<br />

70. Fausnaugh, J.; Kennedy, L.; Regnier, F. Comparison of hydrophobic-interaction and<br />

reversed-phase chromatography of proteins. Journal of Chromatography A 1984, 317, 141-<br />

155.<br />

123


References<br />

71. Xindu, G. and Regnier, F.E. Retention model for proteins in reversed-phase liquid<br />

chromatography. Journal of Chromatography A 1984, 296, 15-30.<br />

72. Dondi, F. and Guiochon, G. Theoretical advancement in chromatography and related<br />

separation techniques. Kluwer Academic Pub: 1992.<br />

73. Guo, D.; Mant, C.T.; Taneja, A.K.; Hodges, R.S. Prediction of peptide retention times in<br />

reversed-phase high-performance liquid chromatography II. Correlation of observed and<br />

predicted peptide retention times factors and influencing the retention times of peptides.<br />

Journal of Chromatography A 1986, 359, 519-532.<br />

74. Guo, D.; Mant, C.T.; Taneja, A.K.; Parker, J.; Rodges, R.S. Prediction of peptide retention<br />

times in reversed-phase high-performance liquid chromatography I. Determination of<br />

retention coefficients of amino acid residues of model synthetic peptides. Journal of<br />

Chromatography A 1986, 359, 499-518.<br />

75. Mant, C.T.; Zhou, N.E.; Hodges, R.S. Correlation of protein retention times in reversedphase<br />

chromatography with polypeptide chain length and hydrophobicity. Journal of<br />

Chromatography A 1989, 476, 363-375.<br />

76. Fountoulakis, M. and Takács, B. Design of protein purification pathways: Application to<br />

the proteome of haemophilus influenzae using heparin chromatography. Protein expression<br />

and purification 1998, 14, 113-119.<br />

77. Leser, E.W. and Asenjo, J.A. Rational design of purification processes for recombinant<br />

proteins. Journal of Chromatography B 1992, 584, 43-57.<br />

78. Ong, S. and Pandey, A. An evaluation of the use of two-dimensional gel electrophoresis in<br />

proteomics. Biomolecular engineering 2001, 18, 195-205.<br />

79. Lin, D.; Tabb, D.L.; Yates, J.R. Large-scale protein identification using mass<br />

spectrometry. Biochimica et Biophysica Acta (BBA)-Proteins & Proteomics 2003, 1646, 1-10.<br />

80. Shevchenko, A.; Henrik Tomas, J.H.; Olsen, J.V.; Mann, M. In-gel digestion for mass<br />

spectrometric characterization of proteins and proteomes. Nature protocols 2007, 1, 2856-<br />

2860.<br />

124


References<br />

81. Shevchenko, A.; Loboda, A.; Ens, W.; Schraven, B.; Standing, K.G.; Shevchenko, A.<br />

Archived polyacrylamide gels as a resource for proteome characterization by mass<br />

spectrometry. Electrophoresis 2001, 22, 1194-1203.<br />

82. YatesIII, J.R. Mass spectrometry: from genomics to proteomics. Trends in Genetics 2000,<br />

16, 5-8.<br />

83. Ekins, S.; Mestres, J.; Testa, B. In silico pharmacology for drug discovery: methods for<br />

virtual ligand screening and profiling. British journal of pharmacology 2007, 152, 9-20.<br />

84. Swaan, P.W. and Ekins, S. Reengineering the pharmaceutical industry by crash-testing<br />

molecules. Drug discovery today 2005, 10, 1191-1200.<br />

85. Wurm, F.M. Production of recombinant protein therapeutics in cultivated mammalian<br />

cells. Nature Biotechnology 2004, 22, 1393-1398.<br />

86. Nfor, B.K.; Verhaert, P.D.E.M.; van der Wielen, L.A.M.; Hubbuch, J.; Ottens, M.<br />

Rational and systematic protein purification process development: the next generation. Trends<br />

in biotechnology 2009, 27, 673-679.<br />

87. Chen, J.; Yang, T.; Cramer, S.M. Prediction of protein retention times in gradient<br />

hydrophobic interaction chromatographic systems. Journal of Chromatography A 2008, 1177,<br />

207-214.<br />

88. Bak, H.; Kyhse-Andersen, J.; Thomas, O.R.T. High-throughput immunoturbidimetric<br />

assays for in-process determination of polyclonal antibody concentration and functionality in<br />

crude samples. Journal of Chromatography B 2007, 848, 142-150.<br />

89. Neuhoff, V.; Arold, N.; Taube, D.; Ehrhardt, W. Improved staining of proteins in<br />

polyacrylamide gels including isoelectric focusing gels with clear background at nanogram<br />

sensitivity using Coomassie Brilliant Blue G-250 and R-250. Electrophoresis 1988, 9, 255-<br />

262.<br />

90. Shevchenko, A.; Tomas, H.; Havlis, J.; Olsen, J.V.; Mann, M. In-gel digestion for mass<br />

spectrometric characterization of proteins and proteomes. Nature protocols 2007, 1, 2856-<br />

2860.<br />

125


References<br />

91. Alonso-Orgaz, S.; Moreno, L.; Macaya, C.; Rico, L.; Mateos-Caiceres, P.J.; Sacristain, D.;<br />

Pacrez-Vizcaano, F.; Segura, A.; Tamargo, J.; Lapez-Farra, A. Proteomic Study of Plasma<br />

from Moderate Hypercholesterolemic Patients. Journal of Proteome Research 2006, 5, 2301-<br />

2308.<br />

92. Cowan, R., and Whittaker, R.G. Hydrophobicity indices for amino acid residues as<br />

determined by high-performance liquid chromatography. ScienceDirect 1990, 3, 75-80.<br />

93. Frishman, D and Argos, P. Knowledge-based protein secondary structure assignment.<br />

Proteins: Structure, Function and Genetics 1995, 23, 566-579.<br />

94. Ciorsac, A.; Craciun, D.; Ostafe, V.; Isvoran, A. Nonlinear correlations in the<br />

hydrophobicity and average flexibility along the glycolytic enzymes sequences. Chaos,<br />

Solitons & Fractals 2011, 44, 191-197.<br />

95. Miyazawa, S., and Jernigan, R. Estimation of effective interresidue contact energies from<br />

protein crystal structures: quasi-chemical approximation. Macromolecules ACS 1985, 18,<br />

534-552.<br />

96. Zimmerman, J.M; Eliezer, N; Simha, R. The characterization of amino acid sequences in<br />

proteins by statistical methods. Journal of Chemical Sciences 1968, 21, 170-201.<br />

97. Grantham, R. Amino Acid difference formula to help explain protein evolution. Science<br />

1974, 185, 862-864.<br />

98. Bhaskaran , R. and Ponnuswamy, P.K. Positional flexibilities of amino acid residues in<br />

globular proteins. International journal of peptide and protein research 1988, 32, 241-255.<br />

99. Tosoh Bioscience Hydrophobic interaction chromatography hand book B09P17A.<br />

100. Regnier, F.E. The role of protein structure in chromatographic behavior. Science 1987,<br />

238, 319-323.<br />

101. Kato, Y.; Kitamura, T.; Hashimoto, T. New resin-based hydrophilic support for highperformance<br />

hydrophobic interaction chromatography. Journal of Chromatography A 1986,<br />

360, 260-265.<br />

102. Ochoa, J.L. Hydrophobic (interaction) chromatography. Biochimie 1978, 60, 1-15.<br />

126


References<br />

103. Lienqueo, M.E.; Mahn, A.; Asenjo, J.A. Mathematical correlations for predicting protein<br />

retention times in hydrophobic interaction chromatography. Journal of Chromatography A<br />

2002, 978, 71-79.<br />

104. Wilkins, M.R.; Gasteiger, E.; Sanchez, J.; Bairoch, A.; Hochstrasser, D.F. Twodimensional<br />

gel electrophoresis for proteome projects: The effects of protein hydrophobicity<br />

and copy number. Electrophoresis 1998, 19, 1501-1505.<br />

105. Shevchenko, A.; Jensen, O.N.; Podtelejnikov, A.V.; Sagliocco, F.; Wilm, M.; Vorm, O.;<br />

Mortensen, P.; Shevchenko, A.; Boucherie, H.; Mann, M. Linking genome and proteome by<br />

mass spectrometry: large-scale identification of yeast proteins from two dimensional gels.<br />

Proceedings of the National Academy of Sciences 1996, 93, 14440.<br />

106. Wilkins, M.R.; Pasquali, C.; Appel, R.D.; Ou, K.; Golaz, O.; Sanchez, J.C.; Yan, J.X.;<br />

Gooley, A.A.; Hughes, G.; Humphery-Smith, I. From proteins to proteomes: Large scale<br />

protein identification by two-dimensional electrophoresis and arnino acid analysis. Nature<br />

Biotechnology 1996, 14, 61-65.<br />

107. Herbert, B.; Righetti, P.; Citterio, A.; Boschetti, E. Sample preparation and<br />

prefractionation techniques for electrophoresis-based proteomics. Proteome Research 2007,<br />

15-40.<br />

108. Fountoulakis, M. and Juranville, J.F. Enrichment of low-abundance brain proteins by<br />

preparative electrophoresis. Analatyical Biochemistry 2003, 313, 267-282.<br />

109. Fountoulakis, M.; Takács, M.; Berndt, P.; Langen, H.; Takács, B. Enrichment of low<br />

abundance proteins of Escherichia coli by hydroxyapatite chromatography. Electrophoresis<br />

1999, 20, 2181-2195.<br />

110. Fountoulakis, M.; Juranville, J.F.; Tsangaris, G.; Suter, L. Fractionation of liver proteins<br />

by preparative electrophoresis. Amino Acids 2004, 26, 27-36.<br />

111. Langen, H.; Takács, B.; Evers, S.; Berndt, P.; Lahm, H.W.; Wipf, B.; Gray, C.;<br />

Fountoulakis, M. Two‐dimensional map of the proteome of Haemophilus influenzae.<br />

Electrophoresis 2000, 21, 411-429.<br />

127


References<br />

112. Gu, Z. and Glatz, C.E. A method for three‐dimensional protein characterization and its<br />

application to a complex plant (corn) extract. Biotechnology and bioengineering 2007, 97,<br />

1158-1169.<br />

113. Fausnaugh, J.L. and Regnier, F.E. Solute and mobile phase contributions to retention in<br />

hydrophobic interaction chromatography of proteins. Journal of Chromatography A 1986,<br />

359, 131-146.<br />

114. Salgado, J.C.; Rapaport, I.; Asenjo, J.A. Prediction of retention times of proteins in<br />

hydrophobic interaction chromatography using only their amino acid composition. Journal of<br />

Chromatography A 2005, 1098, 44-54.<br />

115. Fisher, H.F. A limiting law relating the size and shape of protein molecules to their<br />

composition. Proceedings of the National Academy of Sciences 1964, 51, 1285-1291.<br />

116. Warshel, A. What about protein polarity? Nature 1987, 330, 15-16.<br />

117. Capaldi, R.A. and V, Garret. The low polarity of many membrane proteins. Proceedings<br />

of the National Academy of Sciences 1972, 69, 930-932.<br />

118. Lin, F.Y.; Chen, W.Y.; Hearn, M.T.W. Microcalorimetric studies on the interaction<br />

mechanism between proteins and hydrophobic solid surfaces in hydrophobic interaction<br />

chromatography: effects of salts, hydrophobicity of the sorbent, and structure of the protein.<br />

Analytical Chemistry 2001, 73, 3875-3883.<br />

119. Gekko, K. and Hasegawa, Y. Compressibility-structure relationship of globular proteins.<br />

Biochemistry (N. Y. ) 1986, 25, 6563-6571.<br />

120. Peter A., F. Review: Protein function at thermal extremes: balancing stability and<br />

flexibility. Comparative Biochemistry and Physiology - Part A: Molecular & Integrative<br />

Physiology 2001, 129, 417-431.<br />

121. Aasim, M.; Bibi, N.; Vennapusa, R.R.; Fernández, L. M. Extended DLVO calculations<br />

expose the role of the structural nature of the adsorbent beads during chromatography.<br />

Journal of Separation Science 2012, In Press.<br />

128


References<br />

122. Chen, J.; Luo, Q.; Breneman, C.M.; Cramer, S.M. Classification of protein adsorption<br />

and recovery at low salt conditions in hydrophobic interaction chromatographic systems.<br />

Journal of Chromatography A 2007, 1139, 236-246.<br />

123. Fountoulakis, M. and Takács, B. Design of Protein Purification Pathways: Application to<br />

the Proteome ofHaemophilus influenzaeUsing Heparin Chromatography. Protein expression<br />

and purification 1998, 14, 113-119.<br />

124. Rose, G.D.; Gierasch, L.; Smith, J.A. Turns in peptides and proteins. Advanced protein<br />

chemistry 1985, 37, 109.<br />

125. Miller, S.; Janin, J.; Lesk, A.M.; Chothia, C. Interior and surface of monomeric proteins.<br />

Journal of Molecular Biology 1987, 196, 641-656.<br />

126. Fausnaugh, J.L.; Pfannkoch, E.; Gupta, S.; Regnier, F. High-performance hydrophobic<br />

interaction chromatography of proteins. Analatyical Biochemistry 1984, 137, 464-472.<br />

127. Katti, A.; Maa, Y.; Horváth, C. Protein surface area and retention in hydrophobic<br />

interaction chromatography. Chromatographia 1987, 24, 646-650.<br />

128. Gekko, K. and Hasegawa, Y. Compressibility-structure relationship of globular proteins.<br />

Biochemistry (N. Y. ) 1986, 25, 6563-6571.<br />

129. Jungbauer, A.; Machold, C.; Hahn, R. Hydrophobic interaction chromatography of<br />

proteins: III. Unfolding of proteins upon adsorption. Journal of Chromatography A 2005,<br />

1079, 221-228.<br />

130. Fogle, J.L.; O’Connell, J.P.; Fernandez, E.J. Loading, stationary phase, and salt effects<br />

during hydrophobic interaction chromatography: α-Lactalbumin is stabilized at high loadings.<br />

Journal of Chromatography A 2006, 1121, 209-218.<br />

131. Ueberbacher, R.; Haimer, E.; Hahn, R.; Jungbauer, A. Hydrophobic interaction<br />

chromatography of proteins: V. Quantitative assessment of conformational changes. Journal<br />

of Chromatography A 2008, 1198-1199, 154-163.<br />

129


References<br />

132. Reubsaet, J.L.E. and Vieskar, R. Characterisation of π–π interactions which determine<br />

retention of aromatic compounds in reversed-phase liquid chromatography. Journal of<br />

Chromatography A 1999, 841, 147-154.<br />

133. Vennapusa, R.R.; Tari, C.; Cabrera, R.; Fernandez-Lahore, M. Surface energetics to<br />

assess biomass attachment onto hydrophobic interaction adsorbents in expanded beds.<br />

Biotechnology and Bioprocess Engineering 2009, 43, 16-26.<br />

130

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!