Academia.eduAcademia.edu
Transfer of rol genes and Evaluation of Artemisinin Synthesis in Transgenic Artemisia annua L. and Artemisia dubia Wall. By BUSHRA HAFEEZ KIANI Department of Biochemistry Quaid-i-Azam University Islamabad, Pakistan 2012 Transfer of rol genes and Evaluation of Artemisinin Synthesis in Transgenic Artemisia annua L. and Artemisia dubia Wall. Submitted by BUSHRA HAFEEZ KIANI Thesis Submitted to Department of Biochemistry Quaid-i-Azam University, Islamabad In the partial fulfillment of the requirements for the degree of Doctor of Philosophy in Biochemistry Department of Biochemistry Quaid-i-Azam University Islamabad, Pakistan 2012 In the name of ALLAH who is the Beneficent and The most Merciful, Guardian of faith, The Majestic, The Bestower and The Forgiver, Whose help and guidance I always importune at every step. CERTIFICATE The department of Biochemistry, Quaid-i-Azam University, Islamabad accepts this thesis submitted by Bushra Hafeez Kiani in its present form, as satisfying the thesis requirement for the Degree of Doctor of Philosophy (Ph.D.) in Biochemistry and Molecular Biology. Supervisor: _________________ Prof. Dr. Bushra Mirza External Examiner: _________________ External Examiner: _________________ Chairman: _________________ Prof. Dr. Bushra Mirza Dated: __________________ Declaration I hereby declare that the work presented in this thesis is my own effort except where others acknowledged and that the thesis is my own composition. No part of this thesis has previously been presented for any other degree. Bushra Hafeez Kiani DEDICATION This humble effort is sincerely dedicated to My Abbu Ji whose wishes and devotions Made me capable of achieving such a Success and to my sweet, ever-loving Ammi Ji whose hands always rose for my Success, And My Family (Words are futile to that They have done for me) Contents Acknowledgements i List of Tables iii List of Figures iv List of Abbreviations x Abstract xiii Chapter 1 Introduction and review of Literature 1 1.1 Genus Artemisia, a natural source of artemisinin 2 1.1.1 Evidence of presence of artemisinin in some Artemisia species other than A. annua 2 1.1.2 Evidence of antimalarial activity of some Artemisia species other than A. annua 3 1.1.3 Taxonomic classification of genus Artemisia 3 1.1.4 History of Artemisia 4 1.1.5 Morphology 4 1.1.6 Distribution 6 1.2 Morphology of Artemisia annua 6 1.3 Morphology of Artemisia dubia 8 1.3.1 Selection of Artemisia dubia for artemisinin enhancement 9 1.4 Secondary metabolites and Medicinal Constituents of Artemisia 10 1.5 Medicinal Importance of genus Artemisia 11 1.5.1 Non medicinal importance of the genus Artemisia 11 1.5.2 Toxic effects of Artemisia 12 1.6 Artemisinin 12 1.6.1 Artemisinin Synthesis and Storage 13 1.6.2 Derivatives of artemisinin 14 1.6.3 Mode of action of artemisinin 15 1.6.4 Biological activities of artemisinin 16 1.6.4.1 Malaria 17 1.6.4.2 Schistosomiasis 18 1.6.4.3 Hepatitis B 18 1.6.4.4 Cancer 19 1.6.4.4.1 Effect of artemisinin and its derivatives on different types of cancer 19 1.6.4.5 Herbicides 20 1.7 Improvement of Artemisinin Content 20 1.7.1 Selection of best cultivar 21 1.7.2 Selection of different stages 21 1.7.3 Tissue Culture of Artemisia 22 1.7.4 Transformation 24 1.7.4.1 Agrobacterium-mediated Transformation 24 1.7.4.1.1 Classification of Agrobacterium 24 1.7.4.1.2 Agrobacterium tumefaciens 24 1.7.4.1.3 Agrobacterium rhizogenes 25 1.7.4.2 Molecular mechanism of Agrobacterium tumefacienes mediated transformation 26 1.7.4.2.1 The binary vector strategy 28 1.7.4.2.2 The cointegration strategy 28 1.7.4.3 Production of transformed plants with Ti plasmid 29 1.7.4.4 Mechanism of A. rhizogenes infection 30 1.8 Agrobacterium tumefacienes mediated transformation in Artemisia species 31 1.9 Agrobacterium rhizogenes mediated transformation in Artemisia species 32 1.10 Role of rol genes in Plants 33 1.10.1 Synergistic effects of rol genes 37 1.11 Extraction of artemisinin 38 1.12 Analysis of artemisinin 39 1.13 Biosynthetic pathway of artemisinin 40 1.13.1 Post-IPP terpene biosynthesis 41 1.13.2 Committed steps in artemisinin biosynthesis 42 1.13.3 Regulation of the artemisinin biosynthetic pathway 44 1.14 Aims and objective 46 Chapter 2 Genetic Transformation of Artemisia annua and 47 Artemisia with dubia Agrobacterium rol tumefacienes ABC and genes through Agrobacterium rhizogenes 2.1 Introduction 47 2.2 Materials and methods 48 2.2.1 Glassware and chemicals 48 2.2.1.2 Medium 48 2.2.3 Inoculation area and manipulation tools 48 2.2.4 Culturing of tissues 49 2.2.5 Collection of Plant Material 49 2.2.6 Seed Germination 49 2.2.7 Culture environment 49 2.2.8 Transformation of Artemisia species with rol ABC genes 50 2.2.8.1 Agrobacterium tumefaciens mediated transformation of Artemisia dubia and Artemisia annua with rol ABC Genes 50 2.2.8.2 Plant material 50 2.2.8.3 Bacterial strain and plasmid construction 50 2.2.9 Transformation procedure 51 2.2.9.1 Preparation of Explants 51 2.2.10 Co cultivation 51 2.2.10.1 Co cultivation medium 51 2.2.10.2 Co cultivation 51 2.2.11 Selection and regeneration 52 2.2.12 Rooting 52 2.2.13 Transplantations to pots and acclimatization 53 2.2.14 53 2.2.15 Agrobacterium rhizogenes mediated transformation of Artemisia dubia and Artemisia annua with rol ABC Genes In-vitro plant production and sterilization 2.2.16 2.2.16.1 Media used for establishment of hairy root cultures Preparation of MS shooting medium 54 54 2.2.16.2 Preparation of ½MS rooting medium 54 2.2.16.3 Preparation of nutrient agar medium 54 2.2.16.4 Preparation of B5 solid medium 54 2.2.16.5 Preparation of B5 selection medium 54 2.2.16.6 Preparation of B5 liquid medium 54 2.2.17 Induction of Hairy Roots 55 2.2.17.1 Preparation of A. rhizogenes strain 55 2.2.17.2 55 2.2.17.3 A. rhizogenes infection on Artemisia annua and Artemisia dubia Transfer of hairy root cultures on B5 solid medium 2.2.17.4 Transfer of hairy roots in to B5 liquid medium 56 2.2.18 Molecular analysis 56 2.2.18.1 Isolation of genomic DNA from plant leaves 56 2.2.18.1.1 Composition of CTAB buffer 57 2.2.18.2 Extraction and purification of plasmid DNA 57 53 55 62.2.18.2.1 Solutions 58 2.2.19 Polymerase chain reaction 58 2.2.19.1 Primers used during PCR 59 2.2.19.2 Agarose gel electrophoresis 59 2.2.20 Southern blot analysis 59 2.2.20.1 DNA restriction 60 2.2.20.2 Agarose gel electrophoresis 60 2.2.20.3 Transfer of restriction fragments to membrane 60 2.2.20.4 Labeling of DNA 60 2.2.20.5 Hybridization process 61 2.3 Results 61 2.3.2 Seed Surface Sterilization 61 2.3.3 Medium for Seed Germination 62 2.3.4 Types of explants 65 2.3.5 Transformation 66 2.3.5.1 Effect of Co-cultivation Period on Transformation 66 2.3.5.2 Regeneration of Transgenic Plants 66 2.3.5.2.1 Effect of Antibiotics on Agrobacterium in Regeneration Medium 66 2.3.5.2.2 Regeneration of Transgenic Plants 66 2.3.5.3 Rooting of Transformed Plants 71 2.3.6 Morphological Analysis of Transformed Plants 72 2.3.7 In-vitro plant production and sterilization 74 2.3.8 Production of Hairy Roots 75 2.3.9 Molecular Analysis of Transformed Plants 75 2.3.9.1 Genomic DNA Extraction 75 2.3.9.2 Polymerase Chain Reaction (PCR) 77 2.3.10 Southern Blotting 79 2.3.11 Acclimatization 81 2.4 Conclusion 83 Chapter 3 Analysis of Artemisinin and its Derivatives and their comparison in transformed and un-transformed plants of Artemisia annua and Artemisia dubia 3.1 Introduction 84 3.2 Materials and methods 85 3.2.1 Extraction of artemisinin 85 3.2.2 Reagents used 85 3.2.2.1 Solvents used 85 3.2.2.1.1 Solvent No.1: Mixture of Ethyl acetate and Hexane in the ratio of 5:95 85 86 3.2.2.1.3 Solvent No.2: Mixture of Acetonitrile and Water in the ratio of 50:50 Biomass and Solvent Ratio 3.2.3 Analysis of artemisinin content 86 3.2.3.1 High performance liquid chromatography 86 3.2.3.1.1 Preparation of mobile phase 86 3.2.3.1.2 Preparation of dilutions of standard 87 3.2.3.1.3 Column used for HPLC 87 3.2.3.1.4 Flow of mobile phase through HPLC system 87 3.2.3.1.5 Detector 87 3.2.3.1.6 Retention time of peak 87 3.2.3.1.7 Injection volume 87 3.2.3.1.8 Calibration curve 88 3.2.3.1.9 Quantification of artemisinin and its derivatives in sample 88 3.2.3.1.10 Verification of artemisinin and derivatives of artemisinin 88 3.2.2.1.2 86 concentrations 3.2.4 Statistical analysis 89 3.3 Results 89 3.3.1 Analysis of artemisinin content in transformed and untransformed Artemisia annua and Artemisia dubia 89 3.3.2 Analysis of derivatives of artemisinin in transformed and untransformed Artemisia annua and Artemisia dubia plants 93 concentrations 3.3.3 Statistical Analysis 108 3.3.3.1 Comparative analysis of artemisinin and its derivatives in different tissues of Artemisia annua and Artemisia dubia 108 3.4 Conclusion 109 Chapter 4 Analysis of Metabolic pathway and Trichome development and their comparison in transformed and un-transformed plants of Artemisia annua and Artemisia dubia 4.1 Introduction 4.2 Materials and methods 111 112 112 4.2.1 Analysis of Metabolic Pathway 4.2.1.1 RNA Extraction 112 4.2.1.2 cDNA synthesis 113 4.2.2 Quantitative Real Time Polymerase chain reaction (qRT- 115 PCR) 4.2.3 Analysis of Trichome density 116 4.2.3.1 Environmental Scanning Electron Microscope (ESEM) 117 setup 4.3 Results 117 4.3.1 Analysis of Metabolic pathway and Trichome development 117 4.3.2 Relative expression of genes involved in metabolic pathway 117 of artemisinin biosynthesis 4.3.3 Trichome Development 119 4.3.4 Statistical Analysis 122 4.3.4.1 Analysis of different genes involved in artemisinin 122 production pathway in different tissues of Artemisia annua and Artemisia dubia 4.3.4 Analysis of Trichome density 125 4.4 Conclusion 131 Chapter 5 Analysis of anticancer activity on Breast Cancer Cell lines and comparison in transformed and untransformed Artemisia annua and Artemisia dubia 5.1 Introduction 133 5.2 Materials and methods 134 5.2.1 Analysis of transformed and untransformed A.annua and A.dubia on Breast Cancer Cell lines 134 5.2.2 Sample preparation 134 5.2.2.1 Hexane extraction 134 5.2.2.2 Aqueous extraction 134 5.2.3 Cell Preparation 135 5.2.3.1 Cells 135 5.2.3.2 Cell inoculation 135 5.2.3.3 Fixation protocol 135 5.2.4 Sulforhodamine B (SRB) Antiproliferative Assay 136 5.3 Results 136 5.3.1 Analysis of transformed and untransformed Artemisia 136 annua and Artemisia dubia on Breast Cancer Cell lines. 5.4 Conclusion 143 Chapter 6 Sequence analysis of Artemisia annua and Artemisia dubia and comparison in transformed and untransformed plants 6.1 Introduction 144 6.2 Materials and methods 145 6.2.1 Sequence analysis of transformed and untransformed A.annua and A.dubia 145 6.2.2 Purification and Fragmentation of mRNA 145 6.2.2.1 Formation and purification of RBP 145 6.2.2.2 Formation and purification of RFP 146 6.2.3 Synthesis of First Strand cDNA 148 6.2.3.1 Formation of CDP 148 4.2.4 Synthesis of Second Strand cDNA 149 4.2.4.1 Addition of SSM 149 6.2.5 End Repairing 150 6.2.5.1 Formation of IMP 150 6.2.6 3' Ends Adenylation 151 6.2.6.1 Addition of ATL 152 6.2.7 Ligate Adapters 152 6.2.7.1 Addition of LIG 152 6.2.8 DNA Fragments Enrichment 155 6.2.8.1 PCR reaction 155 6.2.8.2 Amplification of PCR plates 155 6.2.8.3 Cleaning of PCR plates 156 6.2.9 Validation of Libraries 157 6.2.9.1 Quantification of Libraries 157 6.2.9.2 Quality Control 157 6.2.9.2 Evaluation of sequence homology 157 6.3 Results 157 6.4.1 Comparison of various Transgenic Lines with respect to 158 specific Contigs 6.4.2 Contigs of A.annua and A.dubia producing more than 90% 168 homology 6.4.4 Contigs of A.annua and A.dubia producing 80-85% 202 homology Contigs of A.annua and A.dubia do not produce homology 202 6.5 Conclusion 203 Chapter 7 Discussion 7.1 Agrobacterium tumefaciens mediated transformation of 6.4.5 205 Artemisia dubia and Artemisia annua with rol ABC Genes 7.2 Agrobacterium rhizogenes mediated transformation of 210 Artemisia dubia and Artemisia annua with rol ABC Genes 7.3 Analysis of Artemisinin and its Derivatives 212 7.4 Analysis of Metabolic pathway and Trichome development 218 7.5 Analysis of transformed and untransformed Artemisia 220 annua and Artemisia dubia on Breast Cancer Cell lines 7.6 Sequence analysis of Artemisia annua and Artemisia dubia 223 and comparison in transformed and untransformed plants Conclusion 227 Future strategies 228 References 229 Publications 272 Appendices 273 ACKNOWLEDGEMENT I offer my humblest thanks to ALMIGHTY ALLAH the Lord of the creation, the Merciful and the source of all knowledge and wisdom and to His Prophet Hazrat Muhammad (S.A.W) who guided mankind to get out of darkness of illiteracy and ignorance. I am lacking with the words, to express my extreme profound appreciation and sincere thanks to my Supervisor and Chairperson, Department of Biochemistry, Quaid-i-Azam University Islamabad Professor Dr. Bushra Mirza, for her guidance, precious advices, genuine interest, constructive criticism, thoughtful suggestions and generous supervision throughout my research work. This tedious task would not have been possible without her co-operation, kind interest and moral encouragement. I am extremely grateful to Dr. Guy C Barker, Professor and Director of Genomics, University of Warwick, UK for providing research facilities to carry out my research project in his laboratory and also for his cooperation, useful suggestions, inspiration, reassurance, guidance and counseling from time to time to complete this research project. It is my pleasure to mention John S, Irnia N, Linda B, Dr, Raja AA, Dr. Vickey, Dr. Babar M, Dr. Kapila P, Dr Sajjad A, Alison J, Sara T, Anna M, Beyony K, Tina M, Rachel M, for their great cooperation and support during my stay at Wellesbourne, UK. I believe that without their help my Ph.D project may not have been furnished so quickly and nicely. I am also extremely grateful to Isolda R, Ph.D student in Chemistry Department, University of Warwick UK for her kind cooperation during analysis of my samples on Cancer Cell Lines. I am also thankful to Mr. Martin Davis, Research Assitant, Department of Engneering, University of Warwick UK, for his kind cooperation to provide the facility for ESEM. I would like to extend my deepest appreciation to those people, who helped me in one way or another during my stay at Department of Biochemistry, QAU Islamabad, Pakistan. During my PhD research, I worked with a great number of people; I wish to convey my gratitude to all of them specially, Dr. Abdul Mannan, Dr. Ihsan.UL.Haq, Dr. Sheeraz Ahmad, Dr. Waseem Ahmad, Dr. Waheed Ahmad, Dr. Nazif Ullah, Aman Ullah, Waqas Kiani, Tanveer Ahmad, Dr. Naila, Dr. Maryam, Samreen, Laila, Rehana, Nosheen, Samiya, Saira, Maryam and Tooba. I would like to pay cordial thanks to Dr. Naila Safdar for her guidance, Cooperation and care which helped me put up with all the frustrations encountered during my research work. I will always remain thankful to my dearest friends Amama, Fari, Rehana, Rizwana, Humaira, Sheerin, Jasia, Nosheen, Irnia for their moral support and help. I will always remember the cooperation and help of Irshad N, Amir K, Tariq A, and Murad. I am thankful to Higher Education Commission (HEC) Pakistan for providing me indigenous fellowship and IRSIP scholarship. I believe that HEC funding added to my strength a lot. I am also grateful to my Mamu Dr.Amjad Hussain Tariq Kiani and My brother in-law Khaleeq Kiani and Imran Kiani for their encouragement and inspiring guidance in successful completion of this work. I am very thankful to my nephews Nehal, Minaam, Izaan and my nieces Maham and Zoona for their sweet and tension releasing attitude. I will never forget their innocent love for me. My special thanks are reserve for my family for making me what I am today, how can I forget my loving brother, Dr.Qamar Kiani, Sweet Shumaila bhabi, my sweet sisters Noshi baji, Fozi baji and Sobi baji whose prayers are a source of unmatching support for me. Last but not the least, I owe my deepest gratitude to my loving parents whose invaluable prayers and embolding attitude kept my spirit high to strive for the knowledge and integrity which enabled me to reach the milestone (Allah Bless them always, Aameen). I would like to express my apology to those who ever had a soft corner for me but I missed to mention them personally. Finally, all the errors that remains, are mine, alone. Bushra Hafeez Kiani List of Tables No. Title Page no. Table 2.1 Effect of medium on germination of seeds 62 Table 2.2 Morphological differences observed among Transgenic and 73 Control Plants Table 3.1 Analysis of Variance Table for factors affecting Production 108 of Artemisinin and its Derivatives in different tissues of A.annua and A.dubia Table 4.1 Different components used for cDNA Synthesis 114 Table 4.2 Different components used for Master Mix for cDNA 114 Synthesis Table 4.3 Primers used for qRT-PCR 116 Table 4.4 Analysis of Variance Table for factors affecting different 123 genes involved in artemisinin production pathway in different tissues of A.annua and A.dubia Table 4.5 Analysis of Variance for significant difference between 126 different species of Artemisia plant Table 6.1 Comparison of various Transgenic Lines with respect to 159 specific Contigs Table 6.2 Sequences producing Significant Homology 169 List of Figures No Title Page No Fig 1.1 Malaria Distribution and Statistics 1 Fig.1.2 Floral morphology of Artemisia (A) A daisy flower 5 (Asteraceae) – a composite head of many small flowers (B) Detailed description of floral parts. Fig 1.3 Artemisia annua 7 Fig 1.4 Artemisia dubia 8 Fig 1.5 Structure of Artemisinin 12 Fig 1.6 Chemical structure of artemisinin and its derivatives 15 Fig 1.7 Mechanism of the reductive activation of artemisinin by iron 16 (II)-heme, leading to covalent heme-artemisinin adducts. Fig 1.8 Diagramic representation of biological activities of 17 artemisinin Fig 1.9 Structure of the Ti-plasmid of A.tumefacienes 25 Fig 1.10 Structure of the Ri-plasmid of A.rhizogenes 26 Fig 1.11 Transformation of plant cell by Agrobacterium tumefacienes 27 harbouring a wild-type Ti plasmid Fig 1.12 The binary vector strategy. Plasmids A and B complement 28 each other when present together in the same A. tumefaciens cell. The T-DNA carried by plasmid B is transferred to the plant chromosomal DNA by proteins coded by genes carried by plasmid A. Fig 1.13 The cointegration strategy 29 Fig 1.14 Process of hairy roots production through Agrobacterium 31 rhizogenes infection Fig 1.15 Early biosynthetic pathway for artemisinin: steps in the 41 MVA and MEP pathways and post-IPP steps Fig 1.16 Biosynthetic pathway of artemisinin 44 Fig 2.1 Schematic diagram of the T-DNA region of the plasmid 50 No Title Page No pRT99 Fig 2.2 Percentage seed germination with different duration of 63 exposure to 0.1% (w/v) Mercuric Chloride Fig 2.3 In vitro grown seedlings of Artemisia dubia (a, b) 64 and Artemisia annua (c, d) on MS. Fig 2.4 Leaf and Stem Explants used for transformation Artemisia 65 dubia (a, b) and Artemisia annua (c, d) Fig2.5(a) Regeneration response of Artemisia dubia explants on 67 selection media after incubation in Agrobacterium solution for various durations Fig2.5(b) Regeneration response of Artemisia annua explants on selection 68 media after incubation in Agrobacterium solution for various durations Fig 2.6 Effect of incubation period in Agrobacterium solution 69 on regeneration efficiency Fig 2.7 Shooting of transformants from different explants 70 Fig 2.8 Rooting response of transformed plants 72 Artemisia dubia (a, b) and Artemisia annua (c, d) Fig 2.9 Comparison of Transformed and Non-transformed plants 74 Fig 2.10 In-vitro grown plantlet of (a) Artemisia dubia (b) Artemisia 75 annua Fig 2.11 Induction of hairy roots from stems infected with 76 Agrobacterium rhizogenes Fig2.12(a) Amplified product of transgenic Artemisia dubia 78 Fig2.12(b) Amplified product of transgenic Artemisia annua 79 Fig2.13 80 Southern blot analysis (a) Artemisia dubia (b) Artemisia annua Fig 2.14 Transformed plant in green house 82 Fig 3.1 HPLC Chromatograms of (a) Standard Artemisinin (b) 90 Artemisia annua Sample (A2) (c) Spike Sample (A.annua No Title Page No sample + standard Artemisinin) Fig 3.2 HPLC Chromatograms of (a) Standard Artemisinin (b) 91 Artemisia dubia Sample (D2) (c) Spike Sample (A.dubia sample + standard Artemisinin) Fig 3.3 Comparative analysis of Artemisinin content in shoots, roots 92 and hairy roots of transformed and un-transformed plants of Artemisia dubia and Artemisia annua. The average of three plants is shown together with error bars showing SE. A2+Artemisinin and D2+ Artemisinin are the spike samples used for verification of artemisinin concentration. Fig 3.4 HPLC Chromatograms of (a) Standard Artemether (b) 95 Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Artemether) Fig 3.5 HPLC Chromatograms of (a) Standard Artemether (b) 96 Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Artemether) Fig 3.6 HPLC Chromatograms of (a) Standard Arteether (b) 97 Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Arteether) Fig 3.7 HPLC Chromatograms of (a) Standard Arteether (b) 98 Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Arteether) Fig 3.8 HPLC Chromatograms of (a) Standard Dihydroartemisinin 99 (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Dihydroartemisinin) Fig 3.9 HPLC Chromatograms of (a) Standard Dihydroartemisinin 100 (b) Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Dihydroartemisinin) Fig 3.10 HPLC Chromatograms of (a) Standard Artesunate (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Artesunate) 101 No Fig 3.11 Title Page No HPLC Chromatograms of (a) Standard Artesunate (b) 102 Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Artesunate) Fig 3.12 Artemether content in shoots roots and hairy roots of 103 A.annua.A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Artemether and D2+ Artemether are the spike samples used for verification of artemether concentration. Fig 3.13 Arteether content in shoots roots and hairy roots of A.annua. 104 A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Arteether and D2+ Arteether are the spike samples used for verification of arteether concentration. Fig 3.14 Dihydro-artemisinin content in shoots roots and hairy roots 105 of A.annua.A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Dihydroartemisinin and D2+ Dihydroartemisinin are the spike samples used for verification of dihydroartemisinin concentration. Fig 3.15 Artesunate content in shoots roots and hairy roots of A.annua.A1, A2 and A3 represents leaf samples from three 106 No Title Page No different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Arteether and D2+ Arteether are the spike samples used for verification of arteether concentration. Fig 3.16 Comparative and Statistical analysis of Average Artemisinin, 107 Artemether, Arteether, Dihydro-artemisinin and Artesunate content in shoots, roots and hairy roots of transformed and un-transformed plants of A.dubia and A.annua. Each value is the mean of three replicates. Any two means having a common alphabet are not significantly different at p = 0.05 using LSD. Vertical bar represents the standard error of the 3 means. Fig 4.1(A) Expression of genes in leaves of A.annua. 119 Fig 4.1(B) Expression of genes in hairy roots, roots of transformed 120 shoots of A.annua. Fig 4.1(C) Expression of genes in leaves of A.dubia. 121 Fig 4.1(D) Expression of genes in hairy roots, roots of transformed 122 shoots of A.dubia. Fig 4.2 Statistical and comparative analysis of production of 125 artemisinin and its derivatives in different tissues of Artemisia annua and Artemisia dubia. Each value is the mean of three replicates. Any two means having a common alphabet are not significantly different at p = 0.05 using LSD. Vertical bar represents the standard error of the 3 means Fig 4.3(A) Comparative analysis of trichome density in shoots roots and 127 hairy roots of A.annua and A.dubia Fig 4.3(B) Comparison of trichome density in different tissues of transformed and untransformed plants of A.annua and A.dubia 130 No Title Page No Fig 5.1(A) Analysis of percentage survival in hexane fractions of 139 transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia at 50 µg concentration. The IC50 values are shown as values above the bars. *Artemisinin (IC 50 = 0.21 µM) was used as positive control Fig 5.1(B) Analysis of percentage survival in aqueous fractions of 140 transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia at 50 µg concentration. The IC50 values are shown as values above the bars. *Artemisinin (IC 50 = 0.21 µM) was used as positive control Fig 5.1(C) Comparative analysis of average percentage inhibition in hexane and aqueous fractions of transformed 141 and untransformed plants, roots and hairy roots of A.annua and A.dubia at 50 µg concentration. The IC50 values are shown as values above the bars. *Artemisinin (IC 50 = 0.21 µM) was used as positive control Fig 5.2 Comparative correlation analysis between artemisinin content and anticancer activity in hexane and aqueous fractions of transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia. The R2 values are shown as values above the bars. Those samples that had R2 value 0.6-0.9 were significantly correlated. Data was analyzed for significant correlation by using Graph Pad Prism Method. 142 List of Abbreviations ACT Artemisinin-based Combination Therapy ADS Amorpha-4,11-Diene Synthase AIDS Acquired Immune Deficiency Syndrome ANOVA Analysis of Variance APX Ascorbate Peroxidase AsA L-ascorbic Acid ATCC American Type Culture Collection A. Artemisia A. annua Artemisia annua A. radiobacter Agrobacterium radiobacter A. rhizogenes Agrobacterium rhizogenes A. rubi Agrobacterium rubi A. tumefaciens Agrobacterium tumefaciens A. vitis Agrobacterium vitis Bp Base Pair cDNA Complementary Deoxyribonucleic Acid CE-UV Capillary Electrophoresis Coupled to UV Detector Cs Beta-Caryophyllene Synthase CYP Cytochrome P450 CYP71AV1 CYP71AV1 Cytochrome P450 CYP71AV1 DAB 3, 3‟-Diaminobenzidine-HCl DAD Diode Array Detector DEPC Diethylpyrocarbonate DHAA Dihydroartemisinic Acid DMAPP Dimethylallyl Diphophate DMSO Dimethyl Sulfoxide DNA Deoxyribonucleic Acid DNase Deoxyribonuclease DXP 1-Deoxy-D-xylulose-5-Phosphate DXR 1-Deoxy-D-xylulose-5-Phosphate Reductoisomerase. DXS 1-Deoxy-D-xylulose-5-Phosphate Synthase DW Dry Weight ELISA Enzyme-Linked Immunosorbent assay Eps Epi-Cedrol Synthase FDP Farnesyl Diphosphate FPS FDP Synthase Fs (E)-Beta-Farnesene Synthase FW Fresh Weight GC Gas Chromatography G-C Guanine-Cytosine GPX Glutathione Peroxidase GSH Glutathione HMGR Hydroxymethylglutaryl-CoA Reductase HMG-CoA Hydroxymethylglutaryl-CoA HPLC High Performance Liquid Chromatography IPP Isopentenyl Diphosphate L Liter L-ascorbate L-Ascorbic Acid LSD Least Significant Difference mAU Milli Absorbance Units MEP Non-Mevalonate Pathway mL Milli Liter mM Milli Molar mRNA Messenger RNA MS Mass Spectrometry MS medium Murashige and Skoog Medium MVA Mevalonic Acid Pathway NADPH Nicotinamide Adenosine Dinucleotide Phosphate plus Hydrogen NADP Nicotinamide Adenosine Dinucleotide Phosphate NAD Nicotinamide Adenine Dinucleotide NADH Nicotinamide Adenine Dinucleotide plus Hydrogen Ng Nanogram Nm Nanometer PCR Polymerase Chain Reaction RIA Radioimmunoassay Ri plasmids Root Inducing Plasmids RNA Ribonucleic Acid ROS Reactive Oxygen Species rRNA Ribosomal RNA RT-PCR Reverse Transcriptase PCR SOD Superoxide Dismutase SQS Squalene Synthase SQC Sesquiterpene Cyclase T-DNA Transfer DNA TLC Thin Layer Chromatography WHO World Health Organization UV Ultraviolet μg/mL Microgram Per Milliliter ½ MS medium Half Murashige and Skoog medium μM Micro Molar V Volume 18S 18S Ribosomal Small Subunit ΔΔCT Delta-Delta Cycle Threshold H2O2 Hydrogen Peroxide 2,4-D 2,4-Dichlorophenoxyacetic Acid % Percent ºC Degree Centigrade RBP RNA Bead Plate SSM Second Strand Master ATL A-Tailing Mix ALP Adaptor Ligation Plate RFP RNA Fragmentation Plate CDP cDNA Plate ERP End Repair Mix IMP Insert Modification Plate LIG Ligation Mix CAP Clean-up ALP Plate Abstract Abstract Artemisinin is an important secondary metabolite of Artemisia annua and Artemisia dubia. It is a major constituent of Artemisia species. Chemically it is an endoperoxide sesquiterpene lactone. It is a potent antimalarial drug that has also been proven very effective in treatment of cancer. The rol genes have been known to enhance production of secondary metabolites in plants, possibly through stimulation of defense pathway. This study examines the effect of transformation of A.annua and A.dubia with the rol genes through Agrobacterium tumefaciens and Agrobacterium rhizogenes. The artemisinin content, trichome density and expression of key genes in the biosynthetic pathway of artemisinin were measured. Anticancerous activity of extracts of transformed and untransformed A.annua and A.dubia was also observed against MCF-7 breast cancer cell lines. Transcriptomic study of transformed and control A.annua and A. dubia was carried out as well. A number of factors like type of explants, effect of sterilization and cocultivation period have been reported to affect the efficiency of A.tumefacienes mediated transformation. LBA4404 strain of A.tumefacienes containing pRT99 plasmid harboring rol ABC genes were used for the A.tumefacienes mediated transformation. Mercuric chloride 0.1% for 2 minutes showed the best results for seed surface sterilization giving 95% germination. Leaf and stem were found best explants for transformation. Explants were infected with bacterial culture for 5 minutes and cultured on co-cultivation medium (MS medium with 200 µM acetosyringone) for 48 hours. Explants when cultured on selection medium (MS medium containing 0.1mg /L BAP supplemented with 50mg/ml kanamycin), resulted in the maximum number of transformants. Regeneration of transgenic shoots was obtained from both stem and leaf explants on regeneration medium (MS medium containing 0.1mg /L BAP supplemented with 20mg/ml kanamycin and 500mg/l cefotaxime). Eighty percent of the transgenic A.dubia shoots showed rooting response on half MS medium with 0.025mg/L NAA, while transgenic shoots of A.annua produced roots on half MS medium with 0.1mg/L NAA. Control and transgenic plants were transferred to small pots and acclimatized. Morphological differences like increased size and broad leaves were observed. Confirmation of transformation was made through PCR for rol A, B and C genes. Southern blot analysis was performed to check the copy number of inserted genes. A.rhizogenes strain LBA9402 and LBA 8196 carrying rol genes were used for the A.rhizogenes mediated transformation. Transformation with A.rhizogenes was carried out with the plants growing in green house through their in-vitro propagation. Hairy roots were produced from A. rhizogenes strain LBA9402 infected stem portions of A. annua and A.dubia after seven days of infection but no hairy roots were produced from strain LBA8196. Transformed and control roots were cultured on solid B5 medium for further roots proliferation. Transformed roots showed better proliferation than control roots. Artemisinin content was significantly increased in transformed material of both Artemisia species when compared to un-transformed plants. The artemisinin content increased mostly five to ten times within leaves of transformed lines, hairy roots and roots of transformed shoots. It indicated the plant capability of synthesizing much higher amounts than has been achieved so far through traditional breeding. Similarly, amount of different derivatives of artemisinin i.e. artemether, arteether, dihydroartemisinin and artesunate was also significantly increased in transformed material of both Artemisia species when compared to un-transformed plants. Expression of all the tested genes involved in artemisinin biosynthesis pathway was significantly increased, although variation amongst the genes was observed. Cytochrome P450 (CYP71AV1) and aldehyde dehydrogenase 1 (ALDH1) expression levels were higher than that of amorpha-4, 11 diene synthase (ADS). Levels of the trichome development and sesquiterpenoid biosynthetic gene (TFAR1) expression were also found increased in all transgenic lines. Trichome density significantly increased in the leaves of transformed plants, but no trichomes were seen in control or transformed roots. Crude hexane and aqueous extracts of rol genes transformed plants revealed higher anticancerous activities against MCF-7 breast cancer cell lines. Hexane extracts of transgenic plants revealed higher anticancerous activity against MCF-7 breast cancer cell lines compared to aqueous extracts. Transcriptomic study of A.annua and A.dubia allowed sequencing of the transcriptome of these species for the first time. 16400 Contigs were generated by aligning different transcriptome sequences. Up and down regulation and Putative gene functions were predicted. BLAST of 500 contigs was performed, out of which 264 contigs showed homology with genomes of different organisms. Blast results of some contigs showed that some species have genes that are similar to those involved in artemisinin biosynthesis pathway. It would be interesting to know what are the pathways in which those genes are involved in these species. Furthermore, divergence of A. annua and A.dubia from the common ancestors can be found through phylogenetic tree construction. Chapter No. 1 Introduction and Review of Literature INTRODUCTION AND REVIEW OF LITERATURE Malaria is a disease that causes over one million deaths each year, putting 40% of the world population at risk (WHO, 2004) (Fig 1.1). There are four parasites that cause malaria but the one that causes the most illness and death is Plasmodium falciparum. This is because some strains have become resistant to many of the affordable current treatments including, chloroquine, quinine, mefloquine, and primaquine (Balint, 2001). Fig 1.1: Malaria Distribution and Statistics (WHO, 2004) In 1972, China was the first country which purified an important antimalarial drug and discovered its molecular structure. It was named qinghaosu, a sesquiterpene from the plant Artemisia annua L. today it is known as artemisinin (Meshnick et al, 1996). Artemisinin and its derivatives have been found effective against all the stages of resistant strains of P. falciparum (Balint, 2001). Although artemisinin has been found as very useful medicine yet its production is very low as compared to its required amount. The World Health Organization, WHO, estimated that 130 million treatments would be needed in 2010, requiring 330 tons of artemisinin (WHO, 2008). This presents a problem because natural plant produces very low amount of artemisinin. Only 6 kg of artemisinin is produced from one ton of dry A. annua leaves. Researchers are trying to produce an increased amount of artemisinin in different species of Artemisia plants because this drug also cannot be produced economically through organic synthesis (Abdin et al., 2003). 1.1: Genus Artemisia, a natural source of artemisinin Artemisia (Qinghao) is one of the most ancient plants commonly known as ―worm wood‖ that has persistently remained a traditional Chinese herb. It has been described in China as having medicinal properties since 168 BC (Anon, 1982). Ge Hong (281–340 B.C.), in his book ―Handbook of Prescriptions for Emergency Treatment‖, recommended tea-brewed leaves to treat fever and chills. Afterward, antimalarial activity of Artemisia was described in 1596 by Li Shizhen in his book ―Ben Cao Gang Mu‖ (Compendium of Materia Medica) (Efferth, 2009). An antimalarial research program was started by Chinese Government during Vietnam War to search for antimalarial secondary metabolite systematically. As a result, artemisinin (qinghaosu), an active antimalarial secondary metabolite, was isolated in 1972 from A. annua L (Anon, 1981). In 1979, structure of this antimalarial principal was determined (Liu et al., 1979). Artemisinin was a sesquiterpene lactone of terpenes family containing an endoperoxide bridge, named artemisinin or Qinghaosu (Chinese) according to scientific classification of natural products. 1.1.1 Evidence of presence of artemisinin in some Artemisia species other than A. annua Presence of artemisinin in cosmopolitan species, A. annua has been reported in a large number of publications around the globe (Woerdenbag et al., 1990; 1994; Wallaart et al., 2000; Kumar et al., 2002). Although presence of artemisinin in all Artemisia species was suggested by Moore, (1979) yet he had no concrete evidence to support his claim however, further progress in research proved the presence of artemisinin in other species of Artemisia. According to one report artemisinin is present in A. apiacea and A. lancea but only in minor quantities (Hsu, 2006). Zia et al. (2007) described the presence of artemisinin in leaves of A. absinthium. Arab et al. (2006) reported the presence of artemisinin in aerial parts of A. sieberi. Another report also showed the presence of artemisinin in A. dubia and A. indica (Mannan et al., 2008). A. cina is another discovered source of artemisinin (Aryanti et al., 2001). Several naturally growing Artemisia species in Pakistan also contain artemisinin in their various parts (Mannan and Mirza, 2004). 1.1.2: Evidence of antimalarial activity of some Artemisia species other than A. annua There are several reports of antimalarial activity of various species of Artemisia supporting the claims of artemisinin presence in species other than A.annua as well. Zafar et al. (1990) reported antimalarial activity of A. absinthium. Yarnell and Abascal, (2004a; 2004b) also showed the presence of antimalarial activity in extract of A. vulgaris. Valecha et al., (1994) claimed significant antimalarial activity of ethanol and petroleum extracts of A. japonica, A. maritima, and A. nilegaric in rats. Likewise, methanolic extract of A. persica also showed antiplasmodial activity (Sadeghpour et al., 2006). 1.1.3: Taxonomic classification of genus Artemisia The taxonomic classification of genus Artemisia according to Mabberley, (1997) is as follows:  Class: Magnoliopsida  Subclass: Asteridae  Order: Asterales  Family: Asteraceae  Genus: Artemisia Artemisia is one of the largest genera in the family Asteraceae (Martin et al., 2001; 2003; Watson, 2002) and a well-known cosmopolitan wind pollinated genus as well. There are many publications related to its phylogenetic relationships, including its molecular phylogeny (Heywood and Humphries, 1977; Ling, 1982; Watson, 2002; Valles et al., 2003) and distribution (Ling, 1988; 1992; 1994a; 1995; Valles et al., 2003). However, detailed study related to its origin and subsequent development in the geological past, is still lacking. Linnaeus first described this genus as containing 19 species in his Species Plantarum (Linnaeus, 1753). In his later work, most of the species were given their names according to their origin e.g. species reported from Ethiopia was named A. aethiopica L. and that from china as A. chinensis L. Afterward, Hooker and Thomson, (1881) further classified it into sections and placed the species in it according to their annual, biennial or perennial nature. Clapham et al. (1962) recognized seven species through morphological characters, and grouped them into sections. Cullen, (1975) described Artemisia consisting of twenty-three species. There, he focused upon leaf morphology for the identification of species. Later scientists placed Artemisia into various groups like Podlech, (1986) and another group of scientists placed Artemisia in six to eight genera within its taxonomic boundaries. Moreover Artemisia of antiquity was divided into three genera Artemisia, Absinthium and Abrotanum (Cronquist, 1955; 1988; Bremer and Humphries, 1993; Ling, 1994b; Kornkven et al., 1998; 1999; Torrel et al., 1999; Martin et al., 2001). Many local and common names are now used all over the world for a number of eminent Artemisia species. These are Annual wormwood (A. annua), Grand wormwood (A. absinthium), Tarragon (A. dracunculus), Russian Wormwood (A. vestita or A. gmelin) and Mugwort (A. vulgaris). 1.1.4: History of Artemisia Artemisia is a large, assorted group of plants with between 200-400 species belonging to the daisy family Asteraceae and it is well known since ages through the study of taxonomy. It comprises hardy herbs and shrubs known for their volatile oils (Watson, 2002). According to Cullen‘s (1975) view, Artemisia is a late flowering and well pollinated genus. 1.1.5: Morphology Artemisia is annual, biennial or perennial, usually bitter-aromatic or nonaromatic glabrous, hairy or punctuate-glandulase herb or small to large shrub with tap root or horizontal root stock. Leaves alternate, basal large and petiolate, upper often subsessile to sessile, undivided or toothed-shallowly to deeply incised or lobed, palmati or pinnatipartite or 2-4 pinnatisect. Synflorescence a pancicle, sometime racemoid or spicate, occasionally globoid or fascieled capitulla generally numerous, small, rarely mediocre, nodding or erect, oblong-cylindrical or ovate to almost globose, hetorogamacy. Involucre ovoid to campanulate or hemispherical, phyllaries in 2-4 (-6)-sereis imbricate, hairy (or) glabrous inner one mostly scarious-margined or scarious. Receptable plane, convex, conical or hemi-spherical, glabrous or pilose (Fig. 1.2a and 1.2b). Flowers brownish, reddish or yellowish, all tubular. Ray-florets: pistilate and fertile, corolla narrowly tubular, gewrally tapering upwards, 2-3(-4) toothed 1 ± oblique, eligulate, style exerted, 2-cleft, branches recurved, linear filiform and tereteoblong, ± flattened. Discflorets bisexual, fertile or sterile, corolla tubular to ± companulate or funnel shaped, 5 toothed, anthers oblong, larger than the filaments, obtuse or ± cordate at the base, apical appendages lanceolate or subulate, style exserted or included, either 2-cleft with flat, truncate and apically erose or fimbriate, ± recurved branches or columnar with an erose or fimbriate disc or cup at the apex. Cypselas of both marginal and disc (when fertile) terete, oblong-ellipsoid or obvoid to nearly fusiform-prismatic, faintly many striate or smooth, glabrouse, apecially rounded or trunate. Pappus absent or sometimes cypselas with a minute annulus or scarious corolliform ring (Ghafoor, 2002) Figure 1.2: Floral morphology of Artemisia (A) A daisy flower (Asteraceae) – a composite head of many small flowers (B) Detailed description of floral parts. 1.1.6: Distribution Artemisia species are broadly distributed in north temperate regions of both new and old world, extreme west and South Africa, South America (Ghafoor, 2002), Asia, west America, Canada, New Zealand (Stewart, 1972), Afghanistan, and extends westwards to the Atlantic (Said, 1969). Artemisia dubia L. grows wild in northern India, China, Xizang, Nepal and northern hilly areas of Pakistan specially Swat, Parachinar, Abbottabad, Hazara, Bagh (Haq, 1983), Chitral, Srinagar, Pahlgam, Tsrar Sharif, Skardu, Ladakh, Kurram valley, Kishtwarpass (Ghafoor, 2002). Artemisia is highly rich in many areas of Kashmir between altitudes of 5000-7000 ft. (Basu and Kirtikar, 1988; Dasture, 1952). It has also been reported in Margalla hills of Islamabad (Rizwana et al., 2002). Artemisia annua L. is inhabitant to Asia, most probably to China (McVaugh, 1984). A. annua occurs as a part of steppe vegetation in China at 1000 to 1500 meter above sea level (Wang, 1961). It is also commonly found as a wild weed and as a roadside plant. Plant now grows wild in many countries, such as Argentina, Bulgaria, France, Hungry, and Romania and is cultivated for its essential oil in Italy, Spain, Eastern Canada, United States and former Yugoslavia (Said, 1969; Klayman, 1989; 1993). According to Rizwana et al., (2002) 30 species of Artemisia are growing naturally in Pakistan. Another publication indicates the presence of 32 species (Stewart, 1972). However ten-species found in Pakistan are also available in Japan, China and Russia (Junshen et al., 1996). Artemisia is also found in Saudi Arabia (Ghafoor and Al-Turki, 2000). Hajra illustrated 33 species of Artemisia from India (Hajra et al., 1995; Dasture, 1952; Palumin and Stainton, 1992). 1.2: Morphology of Artemisia annua Artemisia annua is an annual herb inhabitant in Asia, especially in China. The name of the plant is qinghao. Argentina, Bulgaria, France, Hungary, Italy, Romania, Spain and many other countries all over the world has naturalized Artemisia annua (Ferreira et al., 1996). A.annua is an annual, biennial, or perennial weed reaching about 2 m in height with alternate branches. Leaves are deeply dissected, with an aromatic odour, 2.5 to 5 cm in length, 1 to 3 cm in width. Flowers are tiny and yellow; in lose panicles with capitula 2 to 3 mm across. There are central and marginal florets. The seed vessels consist of one achaene, faintly nerved and 1 mm long. Naturally the plant is pollinated by insects and by the wind (Hansel et al., 1992) (Fig 1.3). Figure 1.3: Artemisia annua Artemisia annua is annual, biennial, or perennial, usually bitter aromatic or nonaromatic, glabrous, hairy or punctate-glandulose herb or small to large shrubs with taproot or horizontal rootstock. Leaves alternate, basal large and petiolate, upper often sub sessile to sessile, undivided or toothed-shallowly to deeply incised or lobed, palmate- or pinnatipartite or 2-4 pinnatisect. Synflorescence a panicle, sometime racemoid or spicate, occasionally globoid or fascicled. Capitula generally numerous, small, rarely mediocre, nodding or erect, oblong-cylindrical or ovate to almost globose, heterogamous. Involucre ovoid to campanulate or hemispherical: phyllaries in 2-4 (-6)-series, imbricate, hairy or glabrous, inner ones mostly serious-margined or scarious. Receptacle plane, convex, conical or hemispherical, glabrous or pilose. Flowers are brownish, reddish or yellowish, all tubular. Ray-florets: pistilate and fertile; corolla narrowly tubular, generally tapering upward, 2-3(-4)-toothed, ± oblique, eligulate; style exerted, 2-cleft, branches recurved, linear-filiform and tereteoblonge, ± flattened. Disc-florets bisexual, fertile or sterile; corolla tubular to ± campanulate or funnel-shaped, 5-toothed; anthers oblong, longer than the filament, obtuse or ± cordate at the base, apical appendages lanceolate or subulate; style exserted or included, either 2-cleft with flat, truncate and apically erose or fimbriate, ± recurved branches or columnar with an erose or fimbrate disc or cup at the apex. Cypselas of both marginal and discs (when fertile) terete, oblong-ellipsoid or obovoid to nearly fusiform-prismatic, faintly many straite or smooth, glabrous, apically round or truncate. Pappus absent or sometimes cypselas with a minute annulus or scarious corolliform ring (Ghafoor, 2002). 1.3: Morphology of Artemisia dubia Artemisia dubia is a peremial herb with several erect, 1-1.8m tall, sulcate, purpureus glabrescent, staut stems from woody, upright root stock as shown in figure 1.1. Leaves shortly petiolate to almost sessile, oblong-elliptic to broadly ovate, 8-12 x 6-9 cm, green and white dotted above along with or without sparse T-shaped hairs, grayish-green arachnoiad hairy beneath to almost glabrous, bipinnatisect, primary segments, elliptic-lanceolate, 3-4.5 x 1.5-2.5 cm, acute mueronate, secondary segments ± elliptic lanceolate, 1-1.5 x 2.5-4.5 mm, acute or ± obtuse revolute; uppermost leaves linear lanceolate, with or without basal auricles. Capitula numerous, heterogamous, oblong-companulate, 3-3.5 x C.2 mm, ± approximate, almost sessile, in a narrow to broad, 15-30 x 10-20 cm panicle with ascending to ± patent (upper), upto 25 x 2-3 cm primary braneles (Ghafoor, 2002). Involcure 4-seriate, phyllaries loosely imbricate, all sparsely arachnoid hairy outside, outermost narrowly ovate, green, acute, median elliptic to narrowly obovate, 2.5 x 1 mm, obtuse, widely scarious hyaline on margins and apex. Receptacle ± flat, glabrous (Fig 1.4). Figure 1.4: Artemisia dubia Involcure 4-seriate, phyllaries loosely imbricate, all sparsely arachnoid hairy outside, outermost narrowly ovate, green, acute, median elliptic to narrowly obovate, 2.5 x 1 mm, obtuse, widely scarious hyaline on margins and apex. Receptacle ± flat, glabrous. Florets 16-20 purplish, all fertile; marginal florets 6-8 with C. 1mm large, basally broadened, bidentate, glandulose corolla tube and long exerted style branches, disc florets 10-12, with C.1.75mm long, 5 toothed, glandulose apically very sparsely hairy corolla tube. Cypselas brown, 1.25-15 mm large, with terminal carolliform scar (Ghafoor, 2002). Artemisia dubia is a common weed in open localities, fallovo fields, waste places, and roadsides, rare in regularly cultivated fields. Propagate mainly by underground stolens. inhabitant of Europe, cartimental Asia, China, Pakistan, Northern India, and Nepal, used as a medicine in various ways. Host plant for meloid, gyne root knot nematodes. Blooming period is August to October (Ghafoor and AlTyrki, 2002). 1.3.1: Selection of Artemisia dubia for artemisinin enhancement Although artemisinin concentration is highest in Artemisia annua, there are reports describing the presence of artemisinin in Artemisia absinthium, Artemisia dubia and Artemisia indica (Arab et al., 2006; Zia et al., 2007; Mannan et al., 2008). Chemical synthesis of artemisinin is difficult and therefore other sources such as other Artemisia species or tissues such as plant callus (He et al., 1983), shoot (Fulzele et al., 1991) and hairy root cultures (Qin et al., 1994) offer attractive alternatives which could be grown throughout the year distancing production from time of harvest. Mannan et al. (2008) through Agrobacterium rhizogenes transformation increased the artemisinin content within hairy roots of Artemisia dubia and Artemisia indica to 0.6% and to 0.23% respectively. Furthermore, A.annua is restricted to certain global regions due to its growth requirements. A.dubia is capable of growth in more varied conditions and could offer an alternative for production ensuring security of production which would help ensure a constant price. 1.4: Secondary metabolites and medicinal constituents of Artemisia Artemisia is famous due to its distinguished therapeutically significant constituents and has undergone extensive phytochemical investigations over the past two decades. A large number of sesquiterpines have been isolated (Sy and Brown, 1998; 2001) until now from this genus. Most important one is artemisinin (Tu et al., 1982). Artemether, arteether and artesuate are most widely used derivatives of artemisinin which are the methyl ether, ethyl ether and hemisuccinate ester of dihydroartemisinin respectively (Anon, 1982; Li et al., 1982). Artemisinin derivatives have been studied for their efficacy as antimalarial agent. Constituents which have been isolated from its volatile oils (0.5 to 1.6% oil/fresh weight) include α and β thujone (0.042%), phellandrine (12.3%), thygylalcohal (4.5%), azulene (2%), glycoside (0.003%), resin (5.2%), starch (1.45%), alpha-pinene (0.032%), camphene (0.047%), β-pinene (0.882%), myrecene (3.8%),1,8-cineole (5.5%), Artemisia ketone (66.7%), linalool (3.4%), camphor (0.6%), barneole (0.2%), and β-caryophylene (1.2%) (Srivastava, 1999). Other constituents isolated include flavonoids (Hoffmann and Hermann, 1982), ascorbic acid (Slepetys, 1975), carotenoids (Sergeeva and Zakharova, 1977), Tannins (Slepety, 1975), lignans (Greger and Hofer, 1980), Pinitol, artemin, ridentin, santolineal, stigmasterol, daucosteral, sesanin, beta steral, alpha amyrin, Judaicin (cardio tanic), cirsimaritin and glaucolide like sesquiterpine lactones (Tan et al., 1999; Khafagy and Tosson, 1968; Khafagy et al., 1988; Galal et al., 1974). Artemisia plants still represent a large number of unexplored structurally novel compounds that might act as lead for the development of new drugs (Hostettmam, 1987). 1.5: Medicinal importance of genus Artemisia Phytopharmacological evaluation of Artemisia species shows the presence of anti-inflammatory (Sommer et al.,1965), antipyretic (Ikram et al.,1987), antifertility (Rao et al., 1988), antibacterial (Kaul et al.,1976), antifungal (Maruzzella et al., 1960), anthelmintic (Caius and Mahasber, 1920), antiamoebic (Tahir et al., 1991), antimalarial (Hernandez et al., 1990; Zafar et al., 1990), antihepatitic (Gellani and Jambaz, 1995), hepatoprotective (Kiso et al., 1984; Handa et al., 1986; Oshima et al., 1984) and antidiabetic activities (Tan et al., 1999). Studies show that some species of Artemisia (A. judacia L.) have cardiotonic effect (due to Jundaicin) that does not deviate from the general frame work of dagitoxin action (Galal et al., 1974). Antiviral and antibacterial activities are due to the cirsimaritin, which also has inhibitory effect on several mammalian enzymes (Abdalla and Abu-Zarga, 1987). There may be synergism of components too as shown by methoxylated flavenoids of Artemisia dubia, which enhance the activity of artemisinin (Elford et al., 1987). Artemisia annua leaves as well as flowering tops are used medicinally but leaves are preferred. The fresh plant is always more effective than the dry plant. The green tops of this herb are excellent remedy for disorders of the stomach (Dasture, 1952). Traditionally it is used in different cultures as insecticide (Said, 1996) and vermifuge (Caius, 1986). The whole plant is an aromatic tonic and formerly enjoys a high reputation in debility of the digestive organs (Caius, 1986; Basu and Kirtikar, 1988). This species also act as anemma gogue in amenorrhoea, caused by uterine disorders or general debility. It is also prescribed in nocturnal pollutions, chlorosis, anaemia, wasting disease (Dasture, 1952), headaches, migraine, paralysis, facial paralysis, spasmodic affections (such as epiplepsy), and hysteria in nervous irritability nervous disorders and piles. It is an effect diuretic also causes perspiration (Said, 1996). A week decoction of the plant is given to children in measles (Dasture, 1952). Many of its constitutents are used as flavouring agents (Lee and Geismam, 1970; Marco and Barbera, 1990; Heinrich et al., 1998; Sy and Brown, 2001). It is considered effective for jaundice and dropsy. It is used as a tonic either alone or in combinations in atonic diseases of the digestive system (Dasture, 1952). Infusion of top of fresh plants has excellent effect for all disorders of the stomach, creating appetite, promoting digestion and preventing sickness after meals (Caius, 1996). In suitable combinations it is applied as paste over the corresponding area of the abdomen to reduce liver and spleen inflammations (Said, 1969). The herb is commonly prescribed in the form of a poultice or fermentation as an antiseptic and discutiend (Caisus, 1986; Basu and Kirtikar, 1988). Wood tea, or powdered herb in small doses mixed in a little soup is used to relieve bilians melancholia. It is also used in intermittent fever (Caius and Mahaskar, 1920). Indigenous system of medicine uses its leaves and flowering tops in crude form as anthelmintic, antiseptic, febrifuge and stomachic. It has been employed successfully to alleviate chronic fever, dyspepsia and hepatobilary ailments (Said, 1982). Artemisia plant has great potential, much work need to uncover its other activities. That is why; this plant is called ―the plant of future‖. 1.5.1: Non medicinal importance of the genus Artemisia Various constituents obtained from A. pallens (Srivastava, 1999) and A. annua (Charles et al., 1991) have very good fragrance, which can be used in manufacture of high-grade perfumes. A number of specific essential oils obtained from A. annua (Anon, 1992) and A. absinthium (Arino et al., 1999) have wonderful taste that are being used for the flavoring of vermouth and in food industry, respectively. 1.5.2: Toxic effects of Artemisia Numerous species of Artemisia are unhealthy. Periodic uses and large dose of Artemisia causes restlessness, vomiting, vertigo, tremors and canvulsions (Said, 1969). Its affinity to produce headaches and other nervous disorders is well known by travelers in Kashmir and Ladak (Caius, 1986; Basu and Kirtikar, 1988). The juice of the large leaves of Artemisia species is highly nauseous (Dasture, 1952). Many Artemisia species are sources of allergies in humans (Lewis, et al., 1983). Large doses of volatile oils are effective narcotic poison (Caius, 1986). 1.6: Artemisinin Artemisinin an endoperoxide sesquiterpene lactone produced by aerial parts of Artemisia annua L. is effective even against multi-drug resistant strains of the malarial parasite. The isolation and characterisation of artemisinin from Artemisia annua is considered as one of the most novel discoveries in recent medicinal plant research. It was isolated from the plant in 1972 (Roth and Acton, 1989) and in 1979, its structure was determined by X-ray analysis (Brown, 1993). It has an empirical formula of C, H, O. Artemisinin has a peroxide bridge to which its antimalarial properties are attributed. It has a unique structure and lacks nitrogen containing heterocyclic ring, which is found in most anti-malarial compounds as shown in figure 1.4. Artemisinin is an odourless, colourless compound and forms crystals with melting point of 156 - 157 O C. The molecular weight as determined by high resolution mass spectroscopy is m/e 282.1742 m+ (Brown, 1993) (Fig. 1.5). Fig.1.5: Structure of Artemisinin, C15O5H22 (MW=282) Artemisinin has proved to be one of the most promising drugs. It has also shown to possess considerable antimicrobial and antifungal activities (Dhingra et al., 2000). Although the complete organic synthesis has been established, but chemical synthesis of artemisinin is not yet economically feasible because of its complexity and low yield. Currently the leaves, roots and flowers of Artemisia species form the only source of this drug. Artemisinin is found in very low quantities (0.05%-1.1%) in different cultivars of Artemisia dubia and Artemisia annua (1%-4%). High artemisinin yielding clones are being isolated by selection and other non-conventional approaches; however, these have their own limitations. Therefore, in the recent past, in vitro culture system of A.annua has been exploited for the production of artemisinin (Delabays et al., 1993). 1.6.1: Artemisinin synthesis and storage Reports on the distribution of artemisinin throughout the plant are inconsistent. In some studies Artemisinin has been reported to be higher at the top of the plant (Charles et al. 1990; Laughlin 1995) while others suggest it is equally distributed (Laughlin 1995). Artemisinin has also been shown to be produced by differentiated (shoots + roots) shoot cultures (Martinez and Staba 1988; Fulzele et al., 1991; Whipkey et al., 1992; Ferreira and Janick 1996b) but in shoots without roots only trace levels are found suggesting a regulatory effect (Martinez and Staba 1988; Jha et al., 1988; Fulzele et al., 1991; Woerdenbag et al., 1993; Paniego and Giuliette 1994). Most workers (Martinez and Staba 1988; Tawfiq et al., 1989; Fulzele et al., 1991; Kim et al., 1992) were unable to detect artemisinin in roots, although Nair et al., (1986) and Jha et al., (1988) reported trace amounts. Detection of artemisinin from seeds appears to be due to the presence of floral debris (Ferriera et al., 1995b). The highest concentration of artemisinin is found in the inflorescence, which may contain more than ten times as much artemisinin as leaves (Ferriera, et al., 1996). The current understanding is that artemisinin is produced in 10-celled glandular trichomes located on leaves, floral buds, and flowers (Ferreira et al., 1995; Tellez et al., 1999; Olsson et al., 2009) and sequestered in the epicuticular sac at the apex of the trichome (Olsson et al., 2009). For instance, artemisinin concentrations are higher in leaves that are formed later in development than in leaves formed early in the plant‘s development; this difference has been attributed to a higher trichome density and a higher capacity per trichome in the upper leaves (Lommen et al., 2006). Artemisinin biosynthesis was thought not to be synthesized in roots (Ferreira et al., 1995) or pollens. ELISA analysis has shown that green roots accumulate artemisinin (0.001% dry weight) which was confirmed by GC-MS analysis (Jaziri et al., 1995). It has been shown that hairy roots produced by infection with Agrobacterium rhizogenes can produce artemisinin (Jaziri et al., 1995; Liu et al., 1999). Weathers et al., (1994) reported levels of artemisinin (0.4%) comparable to that found in the leaves within hairy root cultures of A. annua transformed with Agrobacterium rhizogenes. This would suggest that the plant is capable of producing artemisinin in the absence of trichomes. However, to date, the regulatory mechanisms that control artemisinin biosynthesis and its formation outside the trichomes are poorly characterized. 1.6.2: Derivatives of artemisinin Derivatives of artemisinin like ether derivatives artemether and arteether, its primary metabolite artenimol and the ester derivative artesunate (Fig 1.6) are exhibiting superior activity against Plasmodium falciparum and also against Plasmodium vivax. The reactivity of endoperoxide bridge which is the common structural feature of artemisinin and all its derivatives is the key for their biological activity. Although artemisinin derivatives are fast acting substances which can rapidly remove the malarial parasites from blood but due to its short biological half-life which preclude its long lasting activity making artemisinin derivatives less preferable in monotherapy, but these derivatives can be used in combination with longer-acting drugs that have a slower onset of activity (Davis et al., 2005). Artemisinin (Qinghaosu) Artenimol (Dihydroartemisinin) Artemether Artesunate Arteether Fig. 1.6: Chemical structure of artemisinin and its derivatives The WHO Roll Back Malaria programme has advocated such a strategy, recommending the use of artemisinin-containing therapies (ACT) in areas of emerging, high resistance to the most commonly used antimalarial (WHO, 2004). Also, a recent publication has reviewed the use of artemisinin-based combination therapies in uncomplicated malaria (Davis et al., 2005). Another asset of artemisinin is their apparently excellent human safety and tolerability (Price et al., 1999). 1.6.3: Mode of action of artemisinin The only one group of compounds for which malarial parasites have not yet developed the resistance are the artemisinin and its derivatives, increasingly being used. It is effective even against chloroquine-resistant and chloroquine sensitive strains of Plasmodium falciparum as well as against cerebral malaria (Anon, 1979; Li et al., 1982; Roth and Acton, 1989; Pras et al., 1991). The unusual structure of artemisinin is the basis of its potent antimalarial activity, which makes it distinctive in mode of action from other antimalarial drugs. There is an assumption that interaction of artemisinin with heme leads to death of the parasite because malarial parasite contains large amount of heme iron (Meshnich et al., 1993; Kamchonwongpaisan and Meshnich, 1996; Meshnich, 1996). Yet, another experiment (Eckstein et al., 2003) showed that artemisinin is activated by ferrous iron inside the parasite, which is independent of heme, the source of this iron is not known. According to Kamchonwongpaisan and Meshnick, (1996) mode of action of artemisinin completes in two steps. First step involves the production of oxygen free radical by the cleavage of the endoperoxide linkage of artemisinin by the heme iron; therefore this step is called as activation step. Oxygen free radical is subsequently rearranged to give a carbon free radical. The second step involves an alkylation of specific malarial proteins by the carbon free radical, which causes a lethal damage to malarial parasite; therefore this step is called an alkylation step (Fig 1.7). Figure 1.7: Mechanism of the reductive activation of artemisinin by iron (II)heme, leading to covalent heme-artemisinin adducts. 1.6.4: Biological activities of artemisinin Artemisinin is not only effective against malaria, but it has also been demonstrated to be effective against a range of other diseases, such as schistosomiasis (Borrmann et al., 2001), hepatitis B (Romero et al., 2005), and a variety of cancer cell lines including breast cancer, human leukemia, colon, and small-cell lung carcinomas (Efferth et al., 2001; Singh and Lai, 2001). Artemisinin is comparatively a safe drug and can be used even for pregnant women without any adverse side effects (Dellicour et al., 2007). With other anti-microbial activities it is also effective against drug resistant cancers (Efferth et al., 2002; Sadava et al., 2002). Artemisinin also act as a natural herbicide because it has been found to be an effective plant inhibitor which contributes to its allelopathic activity (Duke et al., 1987; Chen et al., 1991; Stiles et al., 2005). Entire biological activities of artemisinin are represented in figure 1.8. Figure 1.8: Diagramic representation of biological activities of artemisinin. 1.6.4.1: Malaria A serious disease which is the basis of at least 1 million deaths of children under 5 year, and affected more than 247 million people worldwide and due to which 3.3 billion people were at the risk in 2006 is the disease malaria. According to one report of World health organization (WHO, 2008a), 0.0014 million people were died in 2006 due to 1.5 million malarial successions about which 30% cases were due to the Plasmodium falciparum and one hundred nine countries were affected with malaria in 2008. Malarial parasites which infect human beings are Plasmodium falciparum, Plasmodium vivax, Plasmodium malariae and Plasmodium ovale. Eradication of this disease is becoming very difficult every day due to the development of resistance against multi-drugs system and pesticides by Plasmodium species and the infection vector, the Anopheles mosquito (Winstanley et al., 2002). A large number of publications around the world came in print on the impact of malaria on economic and social development because malaria has a terrific impact on both economic and social development. Shortly, one can agree with the statement that ―where malaria prospers most, human societies have prospered least” (Sachs and Malaney, 2002). 1.6.4.2: Schistosomiasis After malaria, the second most socioeconomically desecrate waterborne disease is bilharziasis, which is commonly known as a schistosomiasis, found in much of the Third World (Warren, 1987). As concern to the socioeconomic and public health, The World Health Organization ranks schistosomiasis as one of the major tropical diseases (WHO, 2008b). Five species Schistosoma mansoni, Schistosoma intercalatum, Schistosoma haematobium, Schistosoma japonicum and Schistosoma mekongi which infect human beings, effects as many as 600 million people as they perform their daily activities related to water such as swimming, fishing, farming, washing, and bathing. In its extremity, schistosomes can survive for decades in their human host, while schistosomiasis is predominantly a disease of the rural poor; travellers to endemic areas are also subject to infection. Although artemisinin is very cogent against this disease but its mode of action is not clear (Garfield, 1986). 1.6.4.3: Hepatitis B All over the world about 2 billion people have been infected with virus and others 350 million people live with incessant infection due to a disease hepatitis B that is 100 times more dangerous than AIDS. According to World Health Organization (WHO, 2008c), every year 0.6 million people die due to acute or chronic effects of hepatitis B, however it is less common in Pakistan (Andre, 2000). Hepatitis B is a viral infection that damages the liver and it can be of two types, acute infections run a course of several months, and chronic infections are often lifelong which lead to liver failure with cirrhosis and hepatocellular carcinoma (Seeger and Mason, 2000). Hepatitis B virus is DNA virus of unique class (Robinson et al., 1974). It is transmitted between people by contact with the blood or other body fluids (i.e. semen and vaginal fluid) of an infected person. Although artemisinin therapy is very effective against this disease but its precise mode of action is unknown (Peterson et al., 1976; WHO, 2008c). 1.6.4.4: Cancer Cancer is a foremost cause of death all over the world. According to different data published all over the world it is described that cancer caused 7.9 million deaths (around 13 % of all deaths) in 2007 and this number is increasing every day with an estimation of 12 million deaths in 2030. The different types of cancer which cause maximum number of deaths every year include Lung, stomach, liver, colon and breast cancer. Cancer cells are produced due to changes in cell acquired by external agents or by inherited genetic factors. According to World Health Organization deaths due to cancer can be prevented up to 30% (WHO, 2008d). To date, the most considerable pharmacological studies on anticancer use of artemisinin have been carried out which explain mechanism of action of artemisinin against cancer in very effective way (Efferth et al., 2001; Singh and Lai, 2001; Efferth et al., 2002; Sadava et al., 2002). According to these considerations, cancer cells require large amount of iron for their multiplication and to express a large concentration of cell surface transferrin receptors that facilitate uptake of the plasma iron-carrying protein transferrin via endocytosis, artemisinin reacts with iron to form free radicals which can kill cells. By covalently tagging artemisinin to transferrin, artemisinin could be selectively picked up and concentrated by cancer cells. In addition, both artemisinin and iron would be transported into the cell in one package. Once an artemisinin-tagged transferrin molecule is endocytosed, iron is released and reacts with tagged artemisinin. As a result, free radicals form that kill the cancer cells. Artemisinin-tagged transferrin is highly selective and potent in killing cancer cells. Thus, artemisinin and artemisinin-tagged iron-carrying compounds could be developed into potent anticancer drugs (Rowen, 2002; Henry et al., 2005). 1.6.4.4.1: Effect of artemisinin and its derivatives on different types of cancer A large number of studies carried out by different laboratories all over the world now explain the effect of artemisinin and its derivatives on different types of cancer, such as breast cancer (Posner et al., 2004; Nam et al., 2007; Sing and Lai, 2004; Thomas et al., 2003), lung cancer (Wu et al., 2006; Thomas et al., 2003), prostate cancer (Chen et al., 2004; Thomas et al., 2003),), head and neck cancer (Yamachika et al., 2004), bladder carcinomas, renal carcinoma (Thomas et al., 2003), ovarian carcinoma (Jiao et al., 2007), cervical carcinoma (Disbrow et al., 2005), pancreas carcinoma (Wu et al., 2001), colon carcinoma (Thomas et al., 2003), thyroid medullary carcinoma (Yamachika et al., 2004), endometrial carcinoma and oral squamous cell carcinoma (Yamachika et al., 2004). The mechanisms of action for their antitumor activities are not fully understood, but may include selective cytotoxicity of cancer cells (Lai and Sing, 1995; Sing and Lai, 2001), induction of apoptosis (Sing and Lai, 2004; Dell‘Eva et al., 2004), modulation of gene expression (Efferth et al., 2002; Wu et al., 2006), causation of cell cycle arrest (Dell‘Eva et al., 2004; Wu et al., 2001; Li et al., 2001) and inhibition of angiogenesis (Dell‘Eva et al., 2004; Wu et al., 2006; Chen et al, 2003; Chen et al., 2004). 1.6.4.5: Herbicides Herbicides are basically chemicals used generally in agriculture, industry, and urban areas for weed control because herbicides inhibit or interrupt normal plant growth and development. According to Chen et al., (1991) artemisinin obtained from A. annua gave same level of inhibition as 2, 4-D and glyphosate in mung bean phaseolus and therefore showed good herbicidal activity against various weeds. Many publications have came into print signifying that artemisinin has potential to be used as natural herbicide (Duke et al., 1987; Chen and Leather, 1990; Yun and Kil, 1992; Lydon et al., 1997) but much effort is needed to find out its mode of action as herbicide. 1.7: Improvement of artemisinin content At present, artemisinin and its derivatives are used in combination therapies for the treatment of malaria (Haynes, 2006), and for treatment of numerous cancers and viral diseases (Efferth, 2007; Efferth et al., 2008). Scientists all over the world are now trying at their best to find out the different ways to increase the production of this precious secondary metabolite to fulfil its increasing demand. Conversely, the major barrier in the commercialization of artemisinin is its low production (0.01-0.8 %) and accumulation in naturally growing Artemisia species (Abdin et al., 2003). At the moment, other resources for its increased production are under investigation, for example the genetic modification of bacteria and yeast, but these experiments at the end only give the artemisinin precursors which need further steps for artemisinin production (Newman et al., 2006; Ro et al., 2006; Shiba et al., 2007; Zeng et al., 2008). Genetic improvement of natural varieties has been attempted but the maximum yield of artemisinin from this route reached so far is 2% (Graham et al., 2010). Due to complexity and high cost in of chemical synthesis (Schmid and Hofheinz, 1983; Haynes 2006; Tue Nguyen, 2011), the biosynthesis of artemisinin via the plant is still the most cost effective. Therefore it is an extensive need to develop and understand the normal biosynthetic pathway of artemisinin to enhance the artemisinin production in Artemisia species and to yield high artemisinin producing transgenic plants and hairy roots for continuous production of artemisinin to meet the increasing demand of this important secondary metabolite to save lives of human beings from many lethal diseases. 1.7.1: Selection of best cultivar Artemisia is a perennial, well-established plant and often woody in nature. Different cultivars of Artemisia are available like Artemisia lactiflora, Artemisia vulgaris, Artemisia frigida, Artemisia absinthium, Artemisia dubia. The only known source of the artemisinin, which is uneconomical to synthesize chemically, is the plant Artemisia annua. The A. annua plant of the family Asteraceae, indigenous to South East Asia, is an annual herb/shrub, which has become naturalized or is in cultivation as a horticultural or medicinal plant in many parts of Asia, Africa, Europe, America and Australia. The natural populations and genetic resources of A. annua from different areas are known to demonstrate considerable variability in the accumulation of artemisinin in the leaves and the capitula of the plant (Charles and Simon, 1990). The foliage and inflorescence have identified and used genotypes of A.annua with high artemisinin content for developing agrotechnologies suitable for cultivating this industrial crop under a variety of soil and agroclimatic conditions (Woerdenbag et al., 1994). Artemisinin yields ranging from 3.2 to 20 kg/h have been reported (Singh et al., 1988; Woerdenbag et al., 1994). 1.7.2: Selection of different stages In the vegetatively growing 20–30 weeks old plants of Artemisia, leaves are the principle organs for the synthesis and accumulation of artemisinin; stems have artemisinin in about 10-fold lower amount. Usually, the younger leaves have more artemisinin than the older leaves. The leaves of young rosette plants during their growth in the winter season (December through March) have very low concentrations of artemisinin. The expression of artemisinin synthesis and accumulation in the leaves progressively increases with the onset of summer in March/April and becomes high by May/June (summer) and peaks during rainy season (July–September).The flowering plants accumulate bulk of their total artemisinin in leaves (30%) and capitula (40%). Since the presence of lipids (oil) in the achenes of capitula makes artemisinin extraction cumbersome, the plants harvested in their vegetative state, when > 90% of the artemisinin is in the leaves and fine stem, offer the best economy in chemical extraction of artemisinin (Sushil et al., 2004; Mannan et al., 2010) Roots lack artemisinin at all stages of plant growth, therefore tissue culture method is adopted to increase the artemisinin content in roots. 1.7.3: Tissue culture of Artemisia Several tissue culture studies have been carried out on various species of Artemisia, findings of some of these are as follows. In one study, round wormwood (Artemisia sphaerocephala Krasch) seeds were germinated on MS medium without plant growth regulators. The hypocotyls of seedlings were sliced and cultured on medium with 2, 4-D to induce callus (Xu and Jia, 1996). Much work has been done on Artemisia annua because of its antimalarial activity. In this species callus can be obtained on media supplemented with combinations of auxins and cytokinins, but non-friable callus is usually obtained. Ferreira and Janick (1996) obtained the highest yield of friable callus with a combination of BA at 4.44 M and 2, 4-D at 4.52 M. Vitamin C reduced browning, and GA3 improved friability. However, after 17 months in culture, only 3 out of 24 clones kept as shoot cultures generated calli. Cell cultures were established using the same callus-inducing medium, without agar. However, artemisinin was not detected from callus or cell cultures or from the liquid medium. Studies have been done to determine the effect of varying strength of B5 culture media, source and level of nitrogen in the media (nitrate and ammonium), phosphate and the phytohormone, gibberellic acid (GA3), on both biomass and artemisinin production in hairy root cultures. Preliminary results show that a 33% increase in B5 medium, as increase in nitrate, an elimination of ammonia, and addition of GA3 increase biomass yields. Artemisinin production was stimulated by regular strength B5 (with lower levels of nitrate than for maximum biomass), low levels of phosphate, and GA3 (Weathers et al., 1996). In another study optimum proliferation of French tarragon (Artemisia dracumulus L. var. sativa) short tips was obtained on MS medium supplemented with 1.8M NAA and 3% sucrose. After 4 weeks of culture, Maximum proliferation was obtained with unpinched shoot tips placed horizontally on the medium. Maximum rooting was observed with cutting>10mm in length. A 5-sec dip of the basal portion of the cuttings in either NAA or IBA increased rooting percentage and root numbers (Mackay and Kitto, 1998). Liu et al., (2002) regenerated new shoots from hypocotyis in Artemisia judaica L. via callus on medium supplemented with TDZ (thiadiazuron). Upto 16 shoots formed per seedling for an exposure of 20 days. Regenerated shoots formed roots when subcultured onto a medium containing indole-3-butyric acid. The regeneration protocol developed in this study provides a basic knowledge for germplasm preservation and for further investigation of medicinally active constituents of A. judaica. Wang and Tan (2002) modified MS medium for enhancing Artemisinin production in Artemisia annua hairy root cultures. They altered the ratio of NO3/NH4+ and the total amount of initial nitrogen. Increasing ammonium to 60mM decreased both growth and Artemisinin accumulation. With NO3-/NH4+ at 5:1 (w/w), the optimum concentration of total initial nitrogen for artemisinin production was 20mM. After 24 days of cultivation with 16.7mM nitrate and 3.3mM ammonium, the maximum artemisinin production of hairy roots was about 14mg/L, a 57% increase over that in the standard MS medium. Nin et al (1996) initiated micropropagated A. annua L. plantlets on MS basal medium supplemented with different concentrations of BA, Kin, NAA, IAA, and 2,4D alone or in combination. Supplementing the medium with low doses of both BAP in combination with NAA, and Kinetin in combination with NAA enhanced the growth rate of callus cultures. Initiation of root and shoot primordial directly from leaf explants cultured on 1.81 M 2,4-D, while adventitious shoot formation from callus was observed occasionally when BA was added to the medium in combination with IAA. Furthermore, medium containing 2.22 M BA and 2.69 M NAA stimulated both callus growth and organogenesis on some callus cultures derived from leaves and stems of young stock material. The best results were obtained with leaf explants. Cytological analysis of root meristems revealed that all regenerants were diploid (2n=18), as expected. 1.7.4: Transformation Transformation technology can be used to introduce genes to produce new phenotypes (Hartwell et al., 2008) or to enhance the expression of existing genes (Royal society, 2002). Effects have been made to enhance the production of artemisinin in Artemisia dubia and Artemisia annua through transformation. The most commonly used method of transformation in plants is Agrobacterium based. 1.7.4.1: Agrobacterium-mediated transformation Agrobacterium tumefaciens has played a major role in the development of plant genetic engineering and the basic research in molecular biology. It accounts for about 80% transgenic plants produced so far. Initially, it was believed that only dicots, gymnosperms and a few monocot species could be transformed by this bacterium, but recent achievements totally changed this view by showing that many ―recalcitrant‖ species not included in its natural host-range such as monocots and fungi can now be transformed (Chan et al., 2004; Bundock et al., 2005). In addition, the transformed cells usually carry single or low copy number T-DNA integrated in their genome with less rearrangement, and very large DNA segments can be transformed into the plants (Liu et al., 1999). 1.7.4.1.1: Classification of Agrobacterium The genus Agrobacterium has been divided into a number of species. However, this division has reflected, for the most part, disease symptomology and host range. The two most commonly used species for transformation purposes are Agrobacterium tumefaciens and Agrobacterium rhizogenes (Stanton and Gelvin, 2003). 1.7.4.1.2: Agrobacterium tumefaciens Agrobacterium tumefaciens is the soil born bacterium which, when containing the Ti plasmid, is able to form crown galls on a number of dicotyledonous plant species (Brown, 2001). Transformation is brought about by the transfer of a large segment of the Ti plasmid called T-DNA to the nuclear genome of a susceptible plant (Chilton et al., 2003). T-DNA contains genes for growth regulator autonomy (Garfinkel et al., 2006 and Joss et al., 1989), and for the synthesis of a wide variety of opines (Murai and Kemp, 1988) which are noble metabolites able to catabolishd by the inciting Agrobacterium (Fig 1.9). Figure 1.9: Structure of the Ti-plasmid of A.tumefacienes (taken from Hooykaas and Schilperoort, 1992) 1.7.4.1.3: Agrobacterium rhizogenes Agrobacterium rhizogenes is a natural plant pathogen responsible for adventitious root formation at the site of infection (Hooykaas, 2004). It incites a disease, hairy root disease, in dicotyledonous plants in a manner very similar to A. tumefaciens (Old and Primrose, 1995). This morphogenic effect of A. rhizogenes is due to the integration and expression of T-DNA of the Ri (root inducing) plasmid in the plant cell genome (Willmitzer et al., 2006). It is also responsible for pathogenicity and the induction of opine synthesis (White and Nester, 2003; Chilton et al., 1982). The neoplastic roots are characterized by a high growth rate and are able to synthesize some secondary metabolites (Flores and Filner, 2007). The Ri plasmid share little homology with Ti-plasmids. The main difference between Ri and Ti plasmid is that transfer of the T-DNA from a Ri plasmid to a plant results not in a crown gall but in hairy root disease, typified by a massive proliferation of a highly branched root system (Brown, 2001) (Fig 1.10). They are of interest because tissue transformed by A. rhizogenes readily regenerates into plantlets, which continue to synthesize opine. For these reasons transformed roots of many medicinal and aromatic plants have been widely studied for the in vitro production of secondary metabolites (Hamill et al., 1986; Benjamin et al., 1993; Jung and Tepfer, 1987; Mano et al., 2005). Figure 1.10: Structure of the Ri-plasmid of A.rhizogenes (taken from Sheela, 2012) Great progress has been made in recent years in studies on the mechanism of Agrobacterium-mediated transformation and its application. Many details of the key molecular events within the bacterial cells involved in T-DNA transfer have been elucidated, and it is notable that some plant factors which were elusive before are purified and characterized (Wei et al., 2000). 1.7.4.2: Molecular mechanism of Agrobacterium tumefacienes mediated transformation The mechanism of T-DNA processing and transfer during Agrobacteirum infection has been subjected to a number of excellent reviews (Knight et al., 2010; Kunitake et al., 2011). Principally, the bacterium can transfer a piece of its plasmid DNA into the infected plant cells, where it integrates into the nuclear genome and expresses its own genes, whose products disrupt the hormonal balance within the plant cells and induce their proliferation to form tumors. In addition, it also produces enzymes to synthesize opines, which the bacteria can use for their own nutrition. The T-DNA is located on a large plasmid called Ti (tumor-inducing)-plasmid, which also contains other functional parts for virulence (vir), conjugation (con) and the origin of its own replication (ori). In the natural infection by wild type bacteria, the T-DNA and the vir genes are essential for inducing plant tumors. The vir region is about 30 kb and encodes at least 10 operons (virA-virJ) whose products are vital to T-DNA processing and transfer. Any genes located in the T-DNA region in principle can be transferred, but they themselves are dispensable for this process. Only the 25-bp direct repeats at the right and the left borders are necessary, of which 14 base pairs are completely conserved and cluster as two separate groups. The infection begins at the wounded sites, and the injured plant cells release some compounds such as the phenolic acetosyringone (AS) acting as specific signals to bind and activate virA, a membrane protein, by autophosphorylation, which subsequently activates VirG by phosphorylating one of its asparate residue. The active form of virG binds specifically to the upstream of other vir genes called vir box, inducing their expression. The virD1 and virD2 are responsible for the T-strand generation, a single-stranded copy of the T-DNA, by specifically recognizing and cutting the bottom strand at the two borders, of which the right one is the start site and thus more important. Figure 1.11: Transformation of plant cell by Agrobacterium tumefacienes harbouring a wild-type Ti plasmid After cutting, virD2 remains covalently attached to the 5' end of the T-strand, forming a complex with, which is then targeted into the nucleus by the nuclear target signals (NLSs) of its associated virD2 and virE2, where the T-DNA randomly integrates into the plant genome as single or multiple copies. VirD2 has an active role in the precise integration of T-strand the plant chromosome (Fig 1.11). Novel strategies have to be developed for inserting new DNA into the plasmid. Two are in general use: 1.7.4.2.1: The binary vector strategy is based on the observation that T-DNA does not need to be physically attached to the rest of the Ti plasmid. A two-plasmid system, with a T-DNA on a relatively small molecule, and the rest of the plasmid in normal form, is just as effective at transforming plant cells. In fact some strains of A. tumefaciens, and related Agrobacteria, have naturally binary plasmid systems. The TDNA plasmid is small enough to have a unique restriction site and to be manipulated using standard techniques (White and Nester, 2003) (Fig 1.12). Figure 1.12: The binary vector strategy. Plasmids A and B complement each other when present together in the same A. tumefaciens cell. The T-DNA carried by plasmid B is transferred to the plant chromosomal DNA by proteins coded by genes carried by plasmid A. 1.7.4.2.2: The cointegration strategy uses as entirely new plasmid based on pBR322 or a similar E. coli vector, but carrying a small portion of the T-DNA. The homology between the new molecule and the Ti plasmid means that if both are present in the same A. tumefaciens cell, recombination can integrate the pBR plasmid into the TDNA region (Fig 1.13). The gene to be cloned is therefore inserted into a unique restriction site on the small pBR plasmid, introduced into A. tumefaciens cells carrying a Ti plasmid, and the natural recombination process integrate the new gene into the T-DNA. Infection of the plant leads to insertion of the new gene, along with the rest of the T-DNA, into the plant chromosome (Brown, 2001). Figure 1.13: The cointegration strategy 1.7.4.3: Production of transformed plants with Ti plasmid As introduction of new genes into every cell in the plant is needed. For this reason a culture of plant cells and protoplasts in liquid medium are infected. A mature plant regenerated from transformed cells will contain the control gene in every cell and will pass the control gene to its offspring. However, regeneration of a transformed plant can occur only if the Ti vector has been ―disarmed‖ so that the transformed cells do not display cancerous properties. Infectivity is mainly controlled by the virulence region of the Ti plasmid i.e., two 25bp repeat sequences found at the left and right borders of the region integrated into the plant DNA. Any DNA placed between these two repeat sequences will be treated as T-DNA and transferred to the plant. A number of disarmed Ti cloning vectors are now available; a typical example is the binary vector pBIN19. The left and right TDNA borders present in this vector flank a copy of the lac Z gene, containing a number of cloning sites, and a kanamycin resistance gene that functions after integration of the vector sequences into the plant chromosome. As with a yeast shuttle vector, the initial manipulations that result in insertion of the gene to be cloned into the pBIN19 are carried out in E. coli, the correct recombinant pBIN19 molecule then being transferred to A. tumefaciens and then into the plant. Transformed plant cells are selected by plating on agar medium containing kanamycin (Brown, 2001). Several Agrobacterium strains and binary vectors have been used for the production of transgenic Artemisia dubia and Artemisia annua (Vergauwe et al., 1998; Nin et al., 2004). 1.7.4.4: Mechanism of A. rhizogenes infection A.rhizogenes agropine type strains carry two T-DNAs, the TL-and the TRDNA (Jouanin, 1988). At least two pathogenic pathways, both leading to root formation by the host, are realized by genes of the TL- and TR-DNA. The TR-DNA harbors two auxin synthesizing genes, aux1 and aux2 (Camilleri and Jouanin, 1999). The auxin biosynthetic pathway catalyzed by the aux1 and aux2 gene products is similar to that in A. tumefaciens and Pseudomonas savastanoi and comprises two steps. First, oxidative decarboxylation of tryptophan leading to indole3-acetamide (IAM) is catalyzed by tryptophan 2-monooxygenase (t2m), encoded by the tms1, aux1, and iaaM genes of A. tumefaciens, A. rhizogenes, and P.savastanoi, respectively (Camilleri and Jouanin 1999). The second step, the conversion of IAM to indole-3-acetic acid (IAA), is catalyzed by IAM hydrolase, the product of the tms2, aux2, and iaaH genes in the above bacteria (Jouanin, 1988). In agropine type A. rhizogenes, the rol ABCD genes are primarily responsible for root induction (White et al., 2008), possibly by enhancing the auxin sensitivity of cells (Maurel et al., 2007; Spano et al., 2006). In some instances, rol B and additional auxin act in concert to induce root formation; the factors alone are not effective individually (Spena et al., 2002; White et al., 2008). TR-DNA genes could provide this additional auxin if both T-DNAs are transferred. However, data indicate that TLDNA located genes in some hosts increase the auxin content, independent of the presence of the TR-DNA. For example, it was shown that TL-DNA transformed pea roots contain increased IAM and IAA concentrations (Prinsen et al., 1992). Interestingly, iaam/tms1, gene expressing, transgenic asparagus and petunia accumulate both IAM and IAA, while tms1-transformed tobacco accumulates only IAM (Prinsen et al., 1998; Van Onkelen et al., 1985). This indicates that conversion of IAM to IAA in cells that express only t2m depends on a host-specific factor such as an endogenous hydrolase (Prinsen et al., 1992). A candidate gene responsible for the increase IAM content in TL-DNA transgenic tissue is the open reading frame 8 (ORF8). Transcripts possibly corresponding to ORF8 have been reported for TL-DNA transformed tissues (Durand- Tardif et al., 2006; Ooms et al., 2008), indicating that the gene is transcribed in infected cells (Fig 1.14). Fig. 1.14: Process of hairy roots production through Agrobacterium rhizogenes infection 1.8: Agrobacterium tumefacienes mediated transformation in Artemisia species Agrobacterium tumefaciens mediated transformation has been used to transform several crops as well as medicinal plants. Biswajit et al. (2000) established transformed organ (petiole, lamina, node and internodes) cultures of the medicinally important Artemisia annua following infection with two wild type nopaline strains of Agrobacterium tumefaciens. Parameters such as explant type, strain type and age of the explants source significantly affected tumorigenesis frequency. Crown galls were formed both on in vivo and in vitro plants: 2-3% of the in vitro galls regenerated spontaneously to produce shooty teratoma of altered phenotype. Artemisinin contents were measured in all transformed as well as non-transformed clones. While shooty teratomas synthesized 0.063 g/100 g DW artemisinin, non-transformed shoots synthesized only 0.0179 g/100 g DW of the compound. A transformation system was developed by Vergauwe et al. (2002a), for Artemisia annua L. plants. Leaf explants from in vitro grown plants developed callus and shoots on medium with 0.05 mg/L NAA and 0.5 mg/L BAP after transformation with the C58C1 RifR (pGV2260) (pTJK136) Agrobacterium tumefaciens strain. A concentration of 20mg/L kanamycin was added in order to select transformed tissue. Kanamycin resistant shoots were rooted on NAA 1.0 mg/L. Polymerase chain reactions and DNA sequencing of the amplification products revealed that 75% of the regenerants contained the foreign genes. 94% of the transgenic plants showed a beta glucuronidase-positive response. In one study, Vergauwe et al. (2005), with a view to optimizing a previously described Agrobacterium tumefaciens-mediated transformation procedure for Artemisia annua, studied the importance of some parameters such as explants type, age of explants source, A. tumefaciens strains and type of binary vector. Several binary vectors were found useful for the production of transgenic callus on explants of different ages. In transformed calli, a good correlation between integration and expression of foreign DNA was observed: Different assays showed expression of beta-glucuronidase, neomycin phosphotransferase II, superoxide dismutase and bleomycin acetyltransferase. The regeneration of transgenic plants required more restricted conditions. Only with the pTJK136 vector could transgenic plants be obtained from leaf and stem explants of 12-18 weeks old plants. Cocultivation of 48 h seemed favourable for the regeneration of transgenic plants. Stable integration and expression of the transgenes were also shown in the progeny. In another study, Mannan et al. (2009) established a transformation method for Artemisia absinthium plants. Different factors such as age and type of explants, presence of ampicillin and kanamycin in Agrobacterium inoculum, concentration of Agrobacterium, infection and co-cultivation time period of Agrobacterium, effects of 2,4-D, pH of co-cultivation medium and effects of cefotaxime in regeneration medium significantly affected transformation of Artemisia absinthium. 1.9: Agrobacterium rhizogenes mediated transformation in Artemisia species Although artemisinin was thought to accumulate only in the aerial part of Artemisia plants (Wallaart et al., 1999), several laboratories have shown that hairy roots can produce artemisinin (Weathers et al., 2004; Jaziri et al., 1995; Liu et al., 1999). Four Asteraceae species were tested: Artemisia dubia, Callendula officinalis, Mikania glomerala and Helianthuus annuus. The explants of all species were inoculated in Agrobacterium rhizogenes strains 8196 and 15834. Artemisia dubia and Artemisia officinalis showed a positive hairy root response. These roots were excised and cultured in vitro. In order to confirm the transgenic character of the hairy roots Southern blot hybridization was carried out (Pellegrino et al., 2008). Transformed cultures of Artemisia annua L. (Asteraceae) were established by the co-culture method using leaf segments of A. annua and Agrobacterium rhizogenes NCIB 8196 or MAFF 03-0172 by Jaziri et al., (1995), the hairy root clones grew vigorously on hormone free medium. The genetic transformation of the root was proved by the opine assay. A highly specific and sensitive enzyme-linked immunosorbent assay (ELISA) method was used for the detection and semi-quantitative determination of artemisinin and structurally related compounds in these cultures. Transformed root cultures of several strains of Artemisia annua were also obtained by infection with Agrobacterium rhizogenes ATCC 15834 and found to be positive for accumulation for artemisinin (Weathers et al., 2004). Naturally, artemisinin accumulates in glandular trichomes, present in leaves, small green stems, flowers buds and seeds of Artemisia (Martinez and Staba, 1988; Ferreira et al., 1997). Although biosynthesis of artemisinin seems to be restricted to the green parts of the plants, and is not synthesized in roots of field-grown plants (Ferreira et al., 1995) or pollens, yet it can be produced by transformed roots of Artemisia dubia and Artemisia indica as described by Mannan et al., (2008). According to Nin et al., (1997), hairy roots were obtained after infection of Artemisia annua shoots with Agrobacterium rhizogenes strains 1855 and LBA 9462. The susceptibility to hairy root transformation varied between plant genotypes and bacterial strains. Hairy roots showed macroscopic differences from control root cultures. Southern blot hybridization confirmed the integration of T-DNA from both p1855 and pBin19, while polymerase chain reaction analysis indicated the presence of the NptII gene in the hairy root genome. Sub-cultured transformed root lines grew well in selective B5 agar-solidified medium containing kanamycin or rifampicin and without hormones. Great differences were found in the profiles of the essential oils isolated from normal and hairy roots. 1.10: Role of rol genes in plants Agrobacterium rhizogenes and Agrobacterium tumefaciens infect wounds of several dicotyledonous and some monocotyledonous species and thereby generate tumorous outgrowths at the site of infection (Nilsson and Olsson, 1997). The genes required for tumorigenesis are found on extra-chromosomal elements, Ti (tumor inducing) plasmids for A. tumefaciens and Ri (root inducing) plasmids for A. rhizogenes, of which a T-DNA (transferred DNA) portion is integrated into the plant genome (Zupan et al., 1995; Zhu et al., 2000). Both pathogens are able to re-specify differentiated cells to gain meristematic functions and, in case of A. rhizogenes, infections are characterized by a massive production of adventitious roots (Meyer et al., 2002). Rol genes have tremendous effect on production of secondary metabolites. Transformation with the rol A gene results in plants with a highly aberrant phenotype, characterized by wrinkled, intensely green leaves, long internodes, dwarfism or semi dwarfism and retarded senescence (Altvorst et al., 1992; Schmulling et al., 1993). The pleiotropic alterations observed in tobacco and potato has led to the hypothesis of a functional imbalance in the auxin/cytokinin ratio in favour of cytokinins (Schmulling et al., 1993). In transgenic tobacco plants a reduction of gibberellic acid content has also been reported, suggesting the involvement of rol A in gibberelin metabolism (Dehio et al., 1998). The rol B protein on the other hand, has been shown to have a tyrosine phosphatase activity and therefore a possible role in the auxin signal transduction pathway (Filippini et al., 1994). Artemisia plants transgenic for rol B have wider leaves and a reduced apical dominance than the wild type (Altvorst et al., 1992). The rol B gene introduced in Nicotiana tabacum was responsible for an auxin like activity, such as enhanced adventitious root formation, which is correlated with increased auxin sensitivity (Maurel et al., 1991; Filippini et al., 1994; Maurel et al., 1994). Similar effect of rol B was also observed in tobacco leaf explants and thin cell layers by Altamura et al. (1998). Hairy roots from different species (Shen et al., 2001; Shen et al., 2003) as well as leaf fragments from hairy root tobacco regenerants (Spano et al., 1988) were shown to be more sensitive to auxin than their normal counterparts. This feature suggests that Ri T-DNA genes induce the proliferation of transformed cells by a unique mechanism, as compared to A. tumefaciens oncogenes, which cause disease by encoding enzymes for hormone biosynthesis (Zambryski et al., 1989). The rol C involved in the released of active cytokinins from their inactive glucosides due to its cytokinin glucosidase activity Estruch et al. (1999). This role is consistent with the observed phenotype of rol C transgenic plants, characterized mainly by a reduction of apical dominance and plant height (Estruck et al., 1999; Nilsson and Olsson, 1997, Schmulling et al., 1993). The integration and expression of the T-DNA of Ti plasmids causes metabolic changes mainly determined by the iaa M and iaa H genes which code for enzymes involved in indoleacetic acid synthesis and the ipt gene whose product catalyses the first step in cytokinin biosynthesis (Yamada et al., 1985). In ornamental plants such as carnation and Petunia the insertion of rol C leads to the expression of advantageous traits, i.e., increased axillary bud break and development, better rooting ability of stem cuttings (Zuker et al., 2001), increased branching and reduction in time to flowering (Winefield et al.,1999). Some data reveals that the rol C gene could affect activity of defence proteins, such as 1, 3-Dglucanases. Expression of the rol C gene in ginseng cells caused significant increase of 1, 3-D-glucanase activity by production of a new enzyme isoform. Activities of other glucanases tested such as 1, 4-D-glucanases and 1, 6-D-glucanase, which are known to be unrelated to defence proteins, were unchanged (Bulgakov et al., 2002). Concerning the rol D gene, its function as an ornithine cyclodeaminase enzyme, catalyzing, catalyzing the conversion of ornithine to proline, has recently been elucidated (Trovato et al., 2001). In transgenic tobacco plants, rol D induces a striking earliness in the induction of the flowering process and an increase in the number of flowers (Mauro et al., 1994) that have been related to an accumulation of proline or to a depletion of ornithine. In Artemisia plants rol D has a pleiotropic effect, affecting traits of economic interest such as plant productivity, as well as characters generally correlated with the defence response to pathogens (Bettini et al., 2003). In this study, the analysis of transgenic plants did not show any morphological modifications. First generation of transgenic plants was found to flower earlier, and showed an increased number of inflorescence and higher fruit yield. They have also shown that the plants harbouring rol D gene were more tolerant to the toxin produced by the fungus fusarium oxysporum in ion leakage experiments, with respect to the untransformed regenerated controls. The organogenetic competence of roots and Agrobacterium rhizogenes induced hairy roots of twelve Artemisia genotypes has been investigated (Peres et al., 2001). Both roots and hairy roots of Artemisia dubia L. derived genotypes regenerated shoots after 24 weeks of incubation on zeatin contained medium. Anatomical analysis showed that shoot regeneration in roots could be direct or indirect, depending on the genotype considered. Hairy roots showed considerable differences in their morphogenetic responses, when compared to the corresponding non-transgenic roots. The differences observed might reflect the influence of the introduced rol genes on hormonal metabolism/sensitivity. Data pointed to the possibility of the use of A. rhizogenes, combined with regenerating Artemisia genotypes, in a very simple protocol, based on genetic capacity instead of special procedures for regeneration, to produce transgenic Artemisia plants expressing rol genes, as well as, genes present in binary vectors. Root proliferation is not due to diffusible cell division factors and a direct interaction of the proteins encoded on the T-DNA with plant hormonal metabolism could not be shown. The T-DNA encodes up to 18 open reading frames, depending on the bacterial strain (Slightom et al., 1986). Insertional mutagenesis showed that insertions in only 4 of the potential 18 loci noticeably affected the morphology of the hairy roots that were produced (White et al., 1985). These loci were named root locus A-D (rol A-d). The rol A, B, C and D genes have been identified as the main determinants of the hairy root disease caused on dicotyledonous plants by the soil bacterium Agrobacterium rhizogenes. The rol A, B, C and D loci correspond, most likely, to open reading frames (ORFs) 10, 11, 12 and 15 of the TL-DNA (Slightom et al., 1986). When individual rol genes are inserted in plants, they have different phenotypic effects that can be at least in part ascribed to modifications in the endogenous hormone equilibrium and in some cases have shown to be potentially interesting also for economic purposes. The roots induced by the integration of T-DNA of various Agrobacterium rhizogenes wild type strains are characterized by an extensive growth, associated with lateral branching, leading to an important mass of adventitious roots exhibiting a typical phenotype. Furthermore, these roots has been found to produce a high yield of secondary metabolites as compared to that of undifferentiated plant cell suspensions, such as indole alkaloid production in Catharantus roseus cultures (Palazon et al., 2003), ginsenoside production in cultures of Catharatus roseus cultures (Palazon et al., 2003), ginsenoside production in cultures of Panax ginseng (Bulgakov et al., 2002), and anthraquinone production in callus cultures of Rubia cordifolia (Bulgakov et al., 2002). The hairy roots obtained from the transformation of Atropa belladonna with A. rhizogenes 15834 produced various amounts of tropane alkaloids, and in most cases higher than in normal field grown plants (Kamada et al., 1986; Saito et al., 2001). The rol A, rol B, and rol C genes together induced root formation and tropane alkaloid biosynthesis in tobacco and in Datura stramonium (Spena et al., 2002; Palazon et al., 1997). The rol A, B and C genes could play a major role in hairy root induction and metabolite production, also leading to an increase in the production of secondary products in A. belladonna hairy roots as is the case with tobacco root cultures (Palazon et al., 1997). The capacity to induce root formation in the host is greatly increased when the rol A and/or C loci are combined with the rol B locus. Root induction is shown to be correlated with the expression of the rol loci. Transgenic plants exhibit all the characteristics of the hairy root syndrome only when all three loci are present and expressed. Although the activity of the rol genes encoded functions is synergistic, each of them appears to independently influence 3 host functions involved in the determination of root differentiation (Spena et al., 2002). 1.10.1: Synergistic effects of rol genes A number of synergistic effects of rol genes have been reported. The presence of rol A, rol B and rol C together dramatically enhances the growth rate and nicotine production, which shows that the effect of these three rol genes was synergistic (Plazon et al., 1997). Several reports suggest that rol A and rol C alter Polyamine (PA) metabolism and plant phenotype (Michael et al., 1998; Sun et al., 1999; MartinTanguy et al., 2006). In another study, transgenic plants containing the rol A constructs showed an increase in the level of PA biosynthetic enzymes and PA titres (Altabella et al., 1995). It appears that the degree of expression of rol A and/or rol C phenotype depends on the level of their penetration, type of tissue used, promoter strength and site of insertion of the foreign gene into the plant genome. Transgenic roots, especially those transformed either by a combination of the three rol genes (A+B+C) or the rol C alone, grew faster than the untransformed roots (Altabella et al., 1995). Putrescine, spermidine and traces of spermine were present in all samples, both in free and bound forms, while rol A roots showed increased levels of free and bound polyamines (Altabella et al., 1995). The higher polyamine contents found in roots transformed by rol A paralleled with higher ornithine and arginine decarboxylase activities as well as higher nicotine production (Altabella et al., 1995). Little is known about mechanisms by which the rol genes interact secondary metabolism. Studying the transgenic for the rol B and rol C genes callus cultures of R. cordifolia, it has been found that anthraquinone (AQ) production was greatly increased in both transformed cultures compared with the non-transformed cultures (Bulgakov et al., 2002). The induction of AQ biosynthesis by the rol genes did not proceed through the activation of the common Ca+2 –dependent NADPH oxidase pathway that mediates signal transduction between an elicitor receptor complex via transcriptional activation of defence genes (Bulgakov et al., 2003). Okadaic acid and cantharidin, inhibitors of protein phosphatases 1 and 2A, caused an increase of AQ production in transgenic cultures. Okadaic acid stimulated AQ accumulation in the non transformed cultures, whereas cantharidin had no effect. These results have shown that different phsophatases are involved in AQ synthesis in normal and transgenic cultures of R. cordifolia (Bulgakov et al., 2003). 1.11: Extraction of artemisinin Conventionally, Chinese extracted the artemisinin from different parts of Artemisia plants with hot water by using common rules of extraction of Chinese medicine (Stoger, 1991). Researchers have developed many methods to extract artemisinin from plant material, on small (Kim et al., 2002; Liu et al., 2007) as well large scale (ElSohly et al., 1990). Extraction can be done from fresh (Filip et al., 2006) as well as dried plant parts (Widmer et al., 2007) with equal efficiency. Ferreira and Janick, (2002) described that artemisinin can be easily extracted with petroleum ether, which boils from 30 to 60°C or other solvent such as chloroform, acetonitrile and ether, which have a boiling point lower than critical temperature for artemisinin stability. Artemisinin can remain stable up to 150°C in neutral solvent as explained by Lin et al. (1985). At the moment, extraction of artemisinin on large scale is carried out by using the methods described by Klayman et al. (1984). This method includes the extraction of artemisinin from dried leaves of Artemisia annua by using petroleum ether and fractionated on a silica column. Another method which is very economical depends upon the use of multiplayer separator and extractor (Acton et al., 1986) but it is only suitable for small-scale extraction. Large amount of artemisinin was extracted by Elsohly et al., (1987), but artemisinic acid tends to elute with artemisinin and the fraction containing artemisinin may require rechromatography to achieve the necessary purity. To resolve this issue, Elsohly et al. (1990) described a practical and economical procedure for the isolation of artemisinin in kilogram quantities by extracting leaves with hexane followed by partitioning with a methyl cyanide phase. This method produced 99% pure artemisinin without any contamination of artemisitene. The common extraction methods for artemisinin neglects artemisinic acid, but an efficient method for extraction of both artemisinic acid and artemisinin from the same material was proposed by Vonwiller et al. (1993) which can then be converted to artemisinin (Xu et al., 1986) that significantly enhanced final yield of artemisinin. However, a ton of dried leaves are needed to produce about 6kg of artemisinin. 1.12: Analysis of artemisinin Artemisinin is an unstable compound and found in very low concentration in plants, therefore its analysis is very complicated because the intact molecule stains poorly, and other compounds interfere in its detection in crude plant extracts. Although artemisinin is very stable even at 150°C but when heated at 180-200°C, degrades into several products (Lin et al., 1985; Luo and Shen, 1987) and it is also sensitive to acid and base treatment (Zheng et al., 1983). Many scientists used thin later chromatography for rapid estimation of artemisinin content in A. annua (Tu et al., 1982; Klayman et al., 1984; Luo and Shen, 1987; Roth and Acton, 1985; Tawfiq et al., 1989; Pras et al., 1991) but this method is not very reliable due to poor staining characteristics of intact molecule and contamination with other constituents of the plant. Many laboratories used HPLC with UV monitoring at 210 nm but presence of constituents that absorb UV at this wavelength completely obliterate the peak of artemisinin (Acton et al., 1985; Zhao and Zeng, 1985; Liersh et al., 1986; Pras et al., 1991), due to this reason some researchers used absorbance range of 220 nm for artemisinin detection (Dhingra et al., 2000). At the moment, many researchers used absorbance UV range of 260 nm in HPLC for artemisinin detection and this is most authentic, economical, practical and extensively used method for the quantification of artemisinin (Smith et al., 1997; Kim et al., 2001; Towler and Weathers, 2007). Reversed phase HPLC is also very useful for the determination of artemisitene in artemisinin, this method can also be used to check the purity level of artemisinin, and for preparative scale purification of these compounds as reported by Eldomiaty et al. (1991). Detection of artemisinin and its derivatives by using HPLC method with polarographic detection has also been used by Zhou and Xu (1988). Variety of methods have been suggested for the analysis of artemisinin; for example Fulzele et al., (1991) and Sipahimalani et al., (1991) used gas chromatographic (GC) method for the analysis of artemisinin at nanograms levels (detection limit 100 ng). Other methods include rapid screening based on mass spectrometry (MS) used for artemisinin related compounds present in crude hexane extract of A. annua (Ranashige et al., 1993). In addition, there are some reports proving the detection of artemisinin in which GC combined with MS (Banthorpe and Brown, 1989; Woerdenbag et al., 1991; 1993) as well as detection by radioimmunoassay or RIA (Zhao et al., 1986) is used. Jaziri et al. (1993) reported the use of enzyme-linked immunosorbent assay (ELISA) method to detect the artemisinin and closely related compounds in crude extract of A. annua. 1.13: Biosynthetic pathway of artemisinin A lot of work has been done to understand biosynthetic pathway of artemisinin. Terpene biosynthesis involves two independent and differently localized mechanisms. Both pathways converge to produce a terpenoid precursor isopentenyl diphosphate (IPP) (Croteau et al., 2000; Lange et al., 2000) (Figure 1.15). The mevalonic acid pathway (MVA) which initiates from acetyl-CoA is located in the cytosol. The key, regulatory, step in this pathway is the conversion of hydroxymethylglutaryl-CoA (HMG) to mevalonate via, a regulatory enzyme, hydroxymethylglutaryl-CoA reductase (HMGR). Several subsequent steps lead to formation of the cytosolic localized pool of IPP (Towler and Weathers, 2007). The other pathway, non-mevalonate pathway (MEP), to IPP begins with pyruvate and occurs in the plastid with no mevalonate intermediate. The first key, regulatory, step is the synthesis of 1-deoxy-D-xylulose-5-phosphate (DXP) via 1-deoxy-Dxylulose-5-phosphate synthase (DXS). DXP is then converted to 2-C-methyl-Derythritol-4-phosphate via 1-deoxy-D-xylulose-5-phosphate reducto-isomerase (DXR). This is the first committed MEP step towards synthesis of terpenes. Several subsequent steps synthesize the plastid pool of IPP (Rohmer et al., 1996). Although IPP present in the cytosol is generally used for the biosynthesis of sterols, sesquiterpenes, triterpenes, and polyterpenes, and the IPP present in the plastid is used for the biosynthesis of monoterpenes, diterpenes and carotenoids (Towler and Weathers, 2007). Translocation of IPP from the cytosol to the plastid or vice versa may occur depending on the needs of the plant (Adam et al., 1998; Lange et al., 2000). Evidence exists suggesting that the two compartmentalized pathways of terpene biosynthesis may communicate with each other to regulate metabolic intermediate availability. There are no absolute restrictions on the compartmentalization of intermediates in the pathway and the degree of separation probably depends on the species and physiological conditions (Hampel et al., 2005). In A. annua, evidence suggests that both the MVA and MEP pathways seem to play a role in artemisinin production. For example, when the cytosolic MVA pathway was disrupted by inhibiting HMGR with mevinolin, artemisinin levels dropped by about 80 %. When DXR in the MEP pathway was inhibited by fosmidomycin, artemisinin levels dropped about 70 %. Use of both inhibitors resulted in no detectable artemisinin production (Towler and Weathers, 2007). 1.13.1: Post-IPP terpene biosynthesis Once IPP is formed and available in the cytosol, the next step towards artemisinin biosynthesis is the production of farnesyl diphosphate (FPP) via farnesyl diphosphate synthase (FPS). Sequence analysis of FPS from A. annua has shown a very close similarity to FPS from other plants (Matsushita et al., 1996). Indeed transgenic A. annua plants expressing FPS under a constitutive promoter produced 34 times more artemisinin than control lines, suggesting that FPS is one of the regulatory points in artemisinin biosynthesis (Chen et al., 2000a). FPP is a branch point, leading to the biosynthesis of triterpenes, polyterpenes, sterols, and sesquiterpenes (Fig 1.15). Figure 1.15: Early biosynthetic pathway for artemisinin: steps in the MVA and MEP pathways and post-IPP steps (taken from Towler and Weathers, 2007) There is evidence suggesting that in plants the option to branch towards either sterols or sesquiterpenes is under coordinate control. To produce sesquiterpenes, a sesquiterpene cyclase (SQC) is the required first catalyst; to produce sterols, the required first catalyst is a specific squalene synthase (SQS). In plants, when sterol production is upregulated, sesquiterpene production is often down regulated, and vice versa. Vogeli and Chappell, (1988) demonstrated that introducing fungal elicitors to tobacco cell culture suspensions caused a rapid increase in sesquiterpenoid production paralleled by a rapid decrease in sterol production. This coordinate control of SQS and SQC genes has also been demonstrated in potato tubers (Yoshioka et al., 1999; Krits et al., 2007). Upon wounding, metabolic flow is directed towards sterols by the up regulation of SQS and downregulation of SQC. On the other hand, exposure of the wound to fungal pathogens or elicitors causes redirection of metabolic flow towards sesquiterpenoid phytoalexins and SQC is upregulated while SQS is down regulated. In A. annua, Towler and Weathers, (2007) showed that coordinate control of these two pathways may be in play. Inhibition of SQS with miconazole caused a significant increase in artemisinin, suggesting that carbon was channelled towards sesquiterpene synthesis once sterol biosynthesis was inhibited. 1.13.2: Committed steps in artemisinin biosynthesis The first committed step in artemisinin biosynthesis (Fig. 1.16) is the cyclization of farnesyl diphosphate (FPP) to generate amorpha-4, 11-diene, catalysed by amorpha-4, 11- diene synthase (ADS) (Mercke et al., 2000; Wallaart et al., 2001). Subsequent oxidation at the C12 position, mediated by the cytochrome P450 enzyme CYP71AV1, leads to artemisinic alcohol (Ro et al., 2006; Teoh et al., 2006). While arteannuin B has been suggested as a late precursor in artemisinin biosynthesis (Sangwan et al., 1993; Zeng et al., 2008), evidence now favours a route from artemisinic alcohol via dihydroartemisinic acid (Bertea et al., 2005; Covello et al., 2007; Covello, 2008). This route is supported by the cloning and characterization of double bond reductase 2 (DBR2), which reduces the D11 (13) double bond of artemisinic aldehyde, but not of arteannuin B (Zhang et al., 2008), and the cloning of aldehyde dehydrogenase 1 (ALDH1), which catalyses the oxidation of artemisinic and dihydroartemisinic aldehyde (Teoh et al., 2009). The conversion of dihydroartemisinic acid to artemisinin, and of artemisinic acid to arteannuin B, has been suggested to occur via enzyme-independent reactions (Sy & Brown, 2002; Brown & Sy, 2004, 2007). Recently, a broad substrate oxidoreductase (RED1) with high affinity for dihydroartemisinic aldehyde and monoterpenes was identified that may have a negative impact on the flux to artemisinin biosynthesis (Ryde´n et al., 2010). IPPI DMAPP FDS (2X IPP+ DMAPP) IPP Farnesyl Diphosphate (FPP) Amorpha-4, 11-dien3 CYP71AV1 Artemisinic Alcohol CYP71AV1 Dihydroartemisinic Alcohol CYP71AV1/ALDHI RED1 DBR2 Artemisinic Aldehyde Dihydroartemisinic Aldehyde CYP71AV1/ ALDHI Artemisinic Acid ALDH1 Dihydroartemisinic Acid Arteannuin B Artemisinin Fig. 1.16: Biosynthetic pathway of artemisinin (taken from Lies et al., 2011) 1.13.3: Regulation of the artemisinin biosynthetic pathway Although many of the genes involved in artemisinin biosynthesis have been isolated and cloned from A. annua, little is known about their regulation. Most of the data relates to the effects of light, culture age, and tissue location on the expression of these genes with most results to date measured in hairy root cultures (Souret et al., 2003; Teoh et al., 2006). The shift from vegetative growth into reproductive growth of A. annua is another factor that has been shown to influence artemisinin production (Ferreira et al., 1995b). Since many terpenoids are floral fragrances, up-regulation of terpene biosynthesis during the shift from vegetative to flowering is not surprising. Previous studies had suggested that a link between flowering and artemisinin biosynthesis exists, although these studies differed in relation to the stage during which peak artemisinin production was reached. One study, for example, showed that peak artemisinin production occurs in the budding stage just before flowering (Liersch et al., 1986; Woerdenbag et al., 1994; Chan et al., 1995), while others reported that peak production was only reached when the flowers were in full bloom (Ferreira et al., 1995b). Some genes related to terpene synthesis in other plants have also been shown to be transcriptionally activated as the shift to flowering occurs (Dudareva et al., 2003). More recently, two separate studies investigated the link between flowering and artemisinin. The flowering promoter factor genes from Arabidopsis, fpf1, and the early flowering gene from Arabidopsis, CONSTANS, were constitutively expressed in A. annua, and although flowering was induced approximately two to three weeks earlier in transgenic lines, there was no corresponding increase in artemisinin biosynthesis. These data suggested that there was no direct regulatory link between flowering and artemisinin synthesis, and that some other factor is likely contributing to the observed increase in artemisinin content as the shift to the reproductive stage progresses (Wang et al., 2004; Wang et al., 2007). Numerous investigations have shown that the Agrobacterium rol genes can induce high levels of secondary metabolites in hairy root cultures of most transformed plant species (Giri and Narasu, 2000; Sevon and Oksman-Caldentey, 2002). We hypothesized that transformation with the rol genes could induce an increase in artemisinin through stimulation of the synthesis pathways. Hence, the aim of the investigation reported here was to compare artemisinin production in transformed plants and hairy roots of Artemisia dubia and Artemisia annua transformed with rol genes by using Agrobacterium tumefaciens and Agrobacterium rhizogenes with non transformed plants to determine their effect. Trichome densities and the genes involved in the synthesis pathway were monitored to ascertain how the rol genes act to influence the production of artemisinin. 1.14: Aims and objectives The following were the main objectives of this study: 1: To optimize and develop the standard transformation procedures by Agrobacterium tumefacienes and Agrobacterium rhizogenes in Artemisia annua and Artemisia dubia. 2: To compare and analyse artemisinin and artemisinin derivatives contents of transformed and non-transformed plants of A. annua and Artemisia dubia. 3: To monitor the metabolic pathway of artemisinin and the genes involved in the synthesis pathway to ascertain how the rol genes act to influence the production of artemisinin in shoots, roots and hairy roots of Artemisia annua and Artemisia dubia. 4: To compare the trichome densities in shoots, roots and hairy roots of transformed and untransformed plants of Artemisia annua and Artemisia dubia and to monitor their effect on artemisinin production. 5: To analyse the anticancerous properties of transformed and non-transformed plants of Artemisia annua and Artemisia dubia. 6: To study the transcriptome of Artemisia annua and Artemisia dubia through sequencing, their comparison in transformed and control plants and to study the homology of their genome sequence. Chapter No. 2 Genetic Transformation of Artemisia annua and Artemisia dubia with rol genes through Agrobacterium tumefacienes and Agrobacterium rhizogenes 2.1: Introduction Artemisinin and its derivatives are used in combination therapies for the treatment of malaria (Haynes, 2006), and for treatment of numerous cancers and viral diseases (Efferth, 2007; Efferth et al., 2008). However, artemisinin and its derivatives are not available to the millions of the world‘s poorest people because of the low yield of artemisinin in naturally growing Artemisia plants (0.1% to 0.5%) of dry weight (Wallaart et al., 1999). The enhanced production of artemisinin, therefore, is highly desirable. Genetic improvement of natural varieties has been attempted but the maximum yield of artemisinin from this route reached so far is 2% (Graham et al., 2010). Through transformation we can enhance the production of artemisinin in Artemisia dubia and Artemisia annua by inserting our desired genes. Plant cells transformed with the rol genes have been observed to show increased production of secondary metabolites. Production of transgenic plants seems to be the most appropriate choice to improve production of secondary metabolites of any plant. For this purpose Agrobacterium mediated transformation method is being used for several medicinal plants including Artemisia species (Nin et al., 1996; Nin et al., 1997; Vergauwe et al., 1998). Optimized conditions may provide a mean to improve the introduction of foreign genes into Artemisia species. Several Agrobacterium strains have been used for the production of transgenic Artemisia annua (Vergauwe et al., 1996; 1998) and Artemisia absinthium (Nin et al., 1997). Vergauwe et al., (1996) developed a transformation system for Artemisia annua by using Agrobacterium tumefaciens strain C58C1 RifR (pGV2260, pTJK136). The aim of present study was to optimize the transformation conditions for A.annua and A.dubia, establishment of hairy root cultures and production of different transgenic lines of both species. Molecular analysis was performed for confirmation of transformation through PCR and Southern blot analysis. A number of factors have been studied like type of explants, different Agrobacterium strains, sterilization conditions and co-cultivation period affecting transformation conditions of Artemisia annua and Artemisia dubia by using Agrobacterium tumefaciens and Agrobacterium rhizogenes. The conditions optimized in this study can be used later to transform other Artemisia species in order to enhance the artemisinin content in plant. 2.2: Materials and Methods The present research work was carried out in Plant Molecular Biology laboratory, Department of Biochemistry, Quaid-i-Azam University, Islamabad and Plant molecular biology lab, department of life sciences, Warwick University, United Kingdom. A brief account of the materials and methods used and the procedures adopted is given below. 2.2.1: Glassware and chemicals Glassware used in all the experiments was made up of borosilicate (Pyrex). All the glassware was cleaned by boiling in a saturated solution of sodium bicarbonate for 1 hour followed by repeated washing in tap water. Thereafter, these were immersed for 30 minutes in 30% HNO3 solution followed by repeated wash in tap water. Washed glassware was further rinsed with distilled water and then dried at 200°C in an oven. Test tubes and flasks were plugged with absorbent cotton. Autoclaving of the glassware was carried out at 121°C, 15 lbs psi for 30 minutes. Chemicals used in all the experiments of tissue culture study were of analytical and molecular biology grade procured from Sigma Chemical Co., USA. Growth regulators and antibiotics (except cefotaxime) were also obtained from Sigma Chemical Co., USA and Melford Laboratories Ltd, UK. Molecular biology products were purchased from Sigma, Invitrogen and Fermentas while kits used were obtained from Qiagen, Promega and MultiTarget Pharmaceuticals. Cefotaxime, sucrose, glucose, gelling agent (gelrite), agar-agar and Hi-Media Bacto-Agar for microbial work were procured from ―DIFCO‖ laboratories, USA. 2.2.1.2: Medium Murashige and Skoog (MS) medium (Murashige and Skoog, 1962) (AppendixIII), (MS; Sigma cat# M5519) was used for germination of seeds in the present investigation. MYA (Appendix-II) and YMB (Appendix-I) media were used for the growth of bacteria. 2.2.3: Inoculation area and manipulation tools Transfer room was cleaned on monthly basis and sprayed with 95% ethyl alcohol. Surgical instruments, Petri dishes, flasks containing distilled water were sterilized in an autoclave at 121 °C and 15psi for 20 minutes. Surgical instruments and Petri dishes were autoclaved in aluminium foil or newspaper. 2.2.4: Culturing of tissues Aseptic transfer of tissues was carried out in a Laminar Flow Cabinet fitted with a HEPA filter. Before using the Laminar Flow Cabinet, working surfaces were swabbed down with 95 % ethyl alcohol. Surgical instruments, Petri dishes, distilled water and culture vessels were brought into the cabinet. Then the working area and instruments were exposed to UV light for 20 minutes to ensure sterility. Tissue culture work was started about half an hour after the UV light was switched off. Surgical instruments were dipped in 95 % ethyl alcohol. After each manipulation, the instruments were again dipped in ethyl alcohol, reflamed and reused. 2.2.5: Collection of plant material The plants and seeds of A.dubia Wall. and A.annua L. were collected from Donga gali Ayubia pipeline (NWFP) and Astour (Northern areas) Pakistan, which were identified by Taxonomy Lab, Department of Plant Sciences, Quaid-i-Azam University, Islamabad, and were stored in refrigerator at 4 ◦C. These seeds keep their vigor for at least 3 years if stored under dry and cool conditions. 2.2.6: Seed germination Stored seeds of Artemisia dubia and Artemisia annua were surface sterilized with 0.1 % (w/v) Mercuric chloride (HgCl2) solution for varying durations of 2, 3, 4 and 6 minutes. Then seeds were rinsed several times with sterile distilled water under laminar flow hood. Seeds were sown under sterile conditions in Petri dishes containing half strength MS salts, supplemented with 3% (w/v) sucrose and solidified with 0.8% (w/v) agar. The pH was adjusted to 5.8 with 1 N NaOH / HCl before the addition of agar. The medium was autoclaved at 15 psi, 121ºC for 20 minutes. This medium is referred to as the Germination Medium (GM). Then these plates were incubated in growth chamber at 25ºC 16 h of photoperiod, illumination of 45 µE m-2 s1 and 60 % relative humidity under aseptic conditions. 2.2.7: Culture environment Tissue cultures were incubated under conditions of well-controlled temperature, light intensity and photoperiod. Temperature of culture room for the present study was maintained at 25  2 ºC with 16 / 8 hours light dark cycle and the light intensity was maintained at 1000 lux throughout the experiments. 2.2.8: TRANSFORMATION OF ARTEMISIA SPECIES WITH ROL ABC GENES 2.2.8.1: Agrobacterium tumefaciens mediated transformation of Artemisia dubia and Artemisia annua with rol ABC Genes The transformation procedure involving tissue culturing was carried out with two different types of explants i.e. Stem and leaf with Agrobacterium tumefacienes strain LBA4404 containing pRT99 harbouring rol ABC genes. 2.2.8.2: Plant material Stem and leaves (0.5-1 cm in length) were cut transversely from 20 days –old in vitro grown seedlings of both Artemisia species and used for genetic transformation. In transformation experiment, 20-40 control explants were used without cocultivation with Agrobacterium. 2.2.8.3: Bacterial strain and plasmid construction Agrobacterium tumefacienes strain LBA4404 harbouring rol ABC genes was kindly provided by Dr. David Tepfer, institute National de la Recherche (INRA), Vesailles, France. T-DNA region of pRT99 carries rol A, rol B and rol C coding sequences, each of which is under the control of CaMV 70S promoter and CaMV 35S terminator. T-DNA of pRT99 also contains NPTII gene with NOS promoter and NOS terminator (Fig 2.1). Fig 2.1: Schematic diagram of the T-DNA region of the plasmid pRT99 RB: Right border; LB: Left border; NOS PRO: Nopaline synthase promoter; NOS TER: Nopaline synthase terminator; 35SPRO: Cauliflower mosaic virus 35S promoter; 35S TER: Cauliflower mosaic virus 35S terminator; 70S PRO: Cauliflower mosaic virus 70S promoter; NPTII: Neomycin phosphortransferase coding region. 2.2.9: TRANSFORMATION PROCEDURE 2.2.9.1: Preparation of explants Leaves and stem segments, were excised from 20 days old seedlings of both Artemisia species and cut into 0.5-1 cm pieces under laminar flow hood and transferred to MS medium containing 200µM acetosyringone in petriplates. These pieces were used as explants for further manipulation. 2.2.10: Co cultivation 2.2.10.1: Co cultivation medium Co-cultivation medium was prepared by adding 200µM acetosyringone to MS, pH was adjusted at 5.8. The medium was prepared in 500ml flasks which were plugged with cotton, covered with aluminum foil and autoclaved at 15psi, 121ºC for 20 minutes. The autoclaved medium was poured under laminar flow hood, into Petri plates (75mm) containing 25ml of medium in each. The media in the Petri plates was allowed to solidify under the sterilized conditions of laminar flow hood. 2.2.10.2: Co cultivation 1. Agrobacterium tumefacienes strain LBA4404 containing pRT99 was grown overnight in 50 ml of liquid MYB medium (0.5% (w/v) yeast extract, 0.05%(w/v) casein hydrolysate , 0.8% (w/v) mannitol, 0.2% (w/v) ammonium sulfate, 0.5%(w/v) sodium chloride, pH 6.6). After inoculation, bacterial cultures were maintained at 28ºC and 120 rev/min in shaking incubator. 2. Bacterial cells from strain were then collected by centrifugation at 3,500 rpm for 10 minutes at 4ºC in 15ml Falcon tubes. 3. The pellets were resuspended in a hormone free MS (Murashige and Skoog, 1962) medium. 4. The density of bacterial strain was adjusted at approximately 5×1010 CFU/ml. 5. After one day of preculturing explants were immersed in the bacterial suspension for 10-15 minutes. 6. Thereafter the explants were blotted on sterilized filter paper and placed on cocultivation medium containing MS medium with 200µM acetosyringone. In each plate about 10 explants were cocultivated. 7. These plates with explants were kept in growth chamber at 27 ºC, 16h of photoperiod, illumination of 45 uEm-² s-¹ and 60% relative humidity, for two days. 2.2.11: Selection and regeneration After three days of cocultivation, the explants were washed with washing medium (WM) that consisted of sterilized MS medium containing 500mg/l cefotaxime to kill bacteria and blotted on sterilized filter paper. All antibiotics were filter sterilized and added to the autoclaved medium. The explants were then transferred to the flasks containing selection medium SM containing MS medium with 0.1mg/l BAP (Benzyl amino purine) before autoclaving, and 20mg/l Kanamycin, 500mg/l cefotaxime was added with the help of micropipette under the laminar flow hood, when the autoclaved media cooled down at 45ºC. These flasks with explants were maintained at 27 ºC, 16h of photoperiod, illumination of 45 µEm-² s-¹ and 60% relative humidity. The explants were then transferred to fresh medium (SM) weekly during the first month. Afterwards subcultures were made every two weeks. After eight weeks, the concentration of cefotaxime was reduced to 50 mg/L. 2.2.12: Rooting For rooting, the developed shoots cut off segments of Artemisia dubia were cultured on Half MS medium containing 0.025mg/L NAA and solidified with 0.1% gelrite while shoots of Artemisia annua were cultured on MS medium containing 0.1mg/L NAA. When roots appeared then sub culturing of parent plants were done in order to increase the number of transformed plants. For sub culturing same medium is used which was used for rooting containing Half MS medium with 0.025mg/L NAA and 0.1mg/L NAA for Artemisia dubia and Artemisia annua respectively, solidified with 0.1% gelrite for solidification. 2.2.13: Transplantations to pots and acclimatization After selection of transformed seedlings, the controlled and kanamycin-resistant plants were transferred individually to small pots containing soil. The soil was made by mixing equal amounts of clay, sand and peat. Small pores were made in their bottoms so that seepage of excessive water may occur at the bottom and may not injure the roots. Plants after transferring to pots were kept enclosed with transparent polythene bags to retain moisture. Acclimatization was done for 15 days in growth room 25±2°C, 16h of photoperiod, illumination of 45 uEm-² s-¹ and 60% relative humidity. Once being hardened, the plants were shifted to green house. Extensive care was taken till they reached maturity. Number of differences was found in morphological features of the control and transgenic antibiotic resistant plants. The morphological parameters were observed and recorded. 2.2.14: Agrobacterium rhizogenes mediated transformation of Artemisia dubia and Artemisia annua with rol Genes Agrobacterium rhizogenes strain LBA9402 and LBA8196 were used for the transformation experiments. A. rhizogenes strains were grown overnight in MYA medium. Plants were infected with the bacterial strains by the following transformation method. 2.2.15: In-vitro plant production and sterilization Transformation with Agrobacterium rhizogenes was carried out with the plants growing in green house through their in-vitro propagation. Small stem (2-4cm) portion with bud of A.dubia and A.annua plants was used. 70% Ethanol (in sterilized water) for 4 min and solution of commercial bleach (20 ml) with sterilized water (80 ml) and tween 20 (100 µl) was used for 15 min with gentle shaking. It was followed by two minutes washing with sterile distilled water for 5 times. New plants appeared from the bud after one month and plantlets were transferred to half MS medium that were later used for transformation. 24 hours incubated Agrobacterium cultures were used for infection by making 2-3 slight cuts in stem portion of A. annua and Artemisia dubia plants with Agrobacterium rinsed scalpel. 2.2.16: Media used for establishment of hairy root cultures Following media were used for establishment and propagation of continuously artemisinin producing hairy root cultures. 2.2.16.1: Preparation of MS shooting medium Shooting medium was prepared by dissolving 4.44 g MS medium (Murashige and Skoog, 1962) (Cat # M404, Phytotechnology Lab) in distilled water along with 3 % D-sucrose (Phytotechnology cat # S391), 2.5 μM benzylaminopurine HCl (Sigma cat # B5920), 0.25 μM naphthaleneacetic acid (Sigma cat # N0640) and 0.5 % g agar gel (Sigma cat # A3301). Distilled water was added to make the final volume 1 litre. Its pH was adjusted to 5.8 and autoclaved before use. 2.2.16.2: Preparation of ½MS rooting medium Rooting medium was prepared by dissolving the 2.22 g MS medium (½ MS medium, Phytotechnology Lab cat # M404) along with 2 % D-Sucrose, (Phytotechnology Cat # S391) and 0.5 % Agar gel (sigma Cat # A3301) in the distilled water and final volume was made up to 1 litre. Its pH was adjusted to 5.8 and autoclaved before use. 2.16.3: Preparation of nutrient agar medium Nutrient agar medium was prepared by dissolving 23 g Nutrient Agar (Sigma Cat # N-9405) in distilled water and the volume was made up to 1 litre. Its pH was adjusted to 7.0 and autoclaved before use. 2.2.16.4: Preparation of B5 solid medium One litre B5 solid medium was prepared by dissolving its 18 nutrients (detail of nutrients is in appendix IV) in freshly prepared distilled water along with 3 % sucrose, 0.25 % phytagel (Sigma cat # P8169) and final volume was made up to 1 litre. Its pH was adjusted to 5.8, autoclaved, poured into petriplates and solidified. 2.2.16.5: Preparation of B5 selection medium One litre B5 selection medium was prepared by adding 600 mg/L of 0.22 μM filtered cefotaxime in autoclaved B5 solid medium. Cefotaxime was mixed thoroughly in B5 solid medium. After mixing, this medium was poured in the petriplates and solidified. 2.2.16.6: Preparation of B5 liquid medium One litre B5 solid medium was prepared by dissolving its 18 nutrients in freshly prepared distilled water along with 3 % sucrose and final volume was made up to 1 litre. Its pH was adjusted to 5.8, autoclaved and 50 ml media transferred in to 250 ml Erlenmeyer flasks. 2.2.17: Induction of hairy roots 2.2.17.1: Preparation of A. rhizogenes strain A. rhizogenes strain LBA9402 and LBA 8196 (Kindly provided by Dr. Fabricio Medina-Bolivar Lab, ABI, and ASU, USA) was used for production of hairy roots. These Agrobacteria were stored at –70 ºC freezer. Pure cultures of Agrobacteria were prepared by streaking on the nutrient agar in the petriplates for 24 hours at 25ºC. For inoculation, single colony of this strain was picked and streak again on the nutrient agar in the petriplates. After inoculation, A.rhizogenes was placed again in the incubator at 25ºC. After 24 hours, A. rhizogenes was ready for infection in stem portion of A. Annua and A.dubia. 2.2.17.2: A. rhizogenes infection on Artemisia annua and Artemisia dubia Twenty eight days old sterile plantlets of A. annua and A.dubia growing in the magenta box were used for infection and transformation. A single colony of A. rhizogenes, 24 hours incubated Agrobacterium cultures, was picked with sterile scalpel which lightly incised into the stem portion of A. annua and Artemisia dubia. In each plantlet, 2-3 incisions were made on different places along the length of stem, by Agrobacterium rinsed scalpel. The infected plantlets were incubated again in growth room having 24 μM m-2 sec-1 cool white fluorescent light and 25ºC with 16/8 light period. After seven days of infection, hairy roots started emerging from Agrobacterium infected stem portions of A. annua and A.dubia. 2.2.17.3: Transfer of hairy root cultures on B5 solid medium These hairy roots were cut from stem portions of both Artemisia species when they were 1-2cm long and transferred to B5 selection medium in petriplates in such a way that one root line was transferred in one plate. These petriplates were placed in dark at 25ºC for two weeks. B5 selection medium was used to kill the excess of A. rhizogenes from hairy roots. After two weeks, 1-2cm long fast growing root tip was excised from these root cultures and transferred to B5 solid medium in the petriplates and incubated again for another two weeks. 2.2.17.4: Transfer of hairy roots in to B5 liquid medium Ten fast growing root tips emerged from single root line grown in petriplates were transferred to 50 ml B5 liquid media in 250 ml Erlenmeyer flasks. These flasks were placed on orbital shaker at 90 rpm in 24 μM m-2 sec-1 cool white fluorescent light and 25ºC. Hairy roots produced and subcultured after every 2 weeks and maintained in liquid medium were used for further analysis. 2.2.18: Molecular analysis For confirmation of the transformation and integration of the desired genes, molecular analysis was carried out through PCR and Southern blot analysis. Plants transformed with Agrobacterium tumefacienes and Agrobacterium rhizogenes strains LBA4404 and LBA8196 respectively were used for PCR and southern blot analysis for rol A, B and C genes. For PCR and southern blot analysis, genomic DNA was isolated from transformed plants, and plasmid DNA was also isolated from the strains of Agrobacterium tumefacienes and Agrobacterium rhizogenes. 2.2.18.1: Isolation of genomic DNA from plant leaves For extraction of genomic DNA from the seedlings simplified CTAB (Cetyl trimethyl ammonium bromide) method of Doyle and Doyle, (1990) was used. DNA was extracted from both transformed and untransformed control plants by grinding individually frozen seedlings to very fine powder using ice cold pestle and mortar. 5ml CTAB extraction buffer at 65°C was added to the powdered leaves and incubated at this temperature for 20 minutes with occasional vigorous shaking. 2ml of chloroform was then added and shake thoroughly and placed on inverter at room temperature for 20 minutes, then centrifuged at 3000 rpm for 5 minutes to resolve phases. After which aqueous phase was transferred to fresh tube and 2ml of isopropanol was added, mixed well and stored on ice for 10 minutes. It was centrifuged again at 3000 rpm for 5 minutes to collect precipitates. The liquid was drained away and sides of tube were dried using blotting paper. To dissolve ppt. 200ul of TE was added. Then 200ul of 4 M LiAc was added and incubated on ice for 20 minutes. It was centrifuged again for 10 minutes at 3000 rpm and to supernatant 1 ml of absolute ethanol was added in a fresh tube and placed on ice for 20 minutes. To collect precipitates the tube was centrifuged again for 10 minutes at 3000 rpm. The liquid was drained away and DNA was dissolved in 200 µl TE by gentle pipeting. 100 µl of 3 M NaOAc was added, and then equal volume of chloroform was added. After which 2 volume of absolute ethanol was added and stored on ice for 5 minutes to precipitate DNA. It was then centrifuged for 10 minutes to collect ppt. Ethanol was evaporated and DNA was dissolved in 200 µl of TE and stored in refrigerator at 20°C. 2.2.18.1.1: COMPOSITION OF CTAB BUFFER Sorbitol 140 mM Tris, Ph 8 220 mM EDTA 22 mM NaCl 800 mM Sarkosyl 1% CTAB 0.8 % These were combined; pH was adjusted to 8 and autoclaved. 2.2.18.2: Extraction and purification of plasmid DNA Minipreparations of plasmid DNA were obtained by alkaline lyses method described below.  Agrobacterium strains were grown overnight in 50ml of liquid MYB medium. Medium was supplemented with 50 mg/L kanamycin sulphate, added to the cold media, after autoclaving under the laminar flow hood. After inoculation, bacterial culture was maintained at 28°C and 120 rev/min in shaking incubator.  1.5 ml of bacterial culture was poured into four microfuge tubes and centrifuged at 14,000 rpm for 15 minutes at 4°C to collect the cells. The supernatant was removed by micropipette and the cells were left to dry.  The pellet was resuspended in 100 µl of solution 1 by vortexing and was left at room temperature for 10 minutes.  200 µl of solution 2 was added and the contents were mixed by inverting the tubes for 4-5 minutes and left at room temperature for 5 minutes.  Then 100 µl of solution 3 was added and the contents were mixed gently for 45 times, the tubes were stored on ice for 20-30 minutes. Low temperature and pH precipitates the denatured DNA which was pelleted by centrifugation at 14,000 rpm for 10 minutes and the supernatant was transferred into fresh tube.  To purify DNA equal volume of chloroform was added and centrifuged again for 5 minutes at 14,000 rpm. The supernatant was transferred to a fresh tube and double volume of ethanol was added to supernatant and stored at -20°C for 1 hour to precipitate DNA and centrifuged for 10 minutes at 14,000 rpm to collect the ppt.  The pellet was washed twice with 70% ethanol to remove any remaining salt. After drying, the pellet was resuspended in 30 µl of TE buffer and was then stored at -20°C in refrigerator. 2.2.18.2.1: SOLUTIONS Solution 1 Solution 2 Glucose 50 mM Tris 25 mM (pH 8.0) EDTA 10 mM (pH 8.0) SDS 1% NaOH 0.2 N Solution 3 Sodium Acetate 3M TE Tris pH 8.0 10 mM EDTA 1 mM 2.2.19: Polymerase chain reaction Polymerase chain reaction for the detection of rol A, B and C gene was performed following the standard method of Taylor, (1991). PCR reaction was performed in 0.2 ml tubes containing 50 µl total reaction mixture. DNA (0.3 µg) was incubated in a final volume of 50 µl with 0.25 µg of reverse and forward primers, 0.2 mM each of d ATP, d GTP, d CTP, and d TTP and 2 mM MgCl2 and 2 units of Taq polymerase and 5 µl of 10X PCR buffer. The reaction mixture was centrifuged for few seconds thorough mixing. The reaction mixture was taken through thermo cycling conditions as: 5 minutes of 95°C for template denaturation followed by 25 cycles of amplification each consisting of 3 steps; 30 seconds at 95°C for DNA denaturation into single strands; one minute at 53-55°C for rol A, 55-57°C for rol B and 54-56°C for rol C primers to hybridize or ―anneal‖ to their complementary sequences on either side of the target sequence; and one minute at 70°C for extension of complementary DNA strand from each primer. Final 10 minutes at 70°C for Taq polymerase to synthesize any unextended strands left. PCR was performed using gene Amp PCR system 2400 and gene Amp PCR system 9600 thermocyclers (Perkin Elmer, USA). 2.2.19.1: Primers used during PCR During PCR following forward and reverse primers for rol A, B and C gene were used. 1. rol A gene forward 5‘-AGAATGGAATTAGCCG GACTA-3‘, reverse primer 5‘-GTATTAATCCCGTAGGTTTGTT-3‘, 2. rol B forward primer 5‘-GCTCTTGCAGTGCTAGATTT-3‘, reverse primer 5‘-GAAGGTGC AAGCTACCT CTC-3‘, 3. rol C gene forward 5‘-GAAGACGACCTGTGTTCTC-3‘, reverse primer 5‘CGTTCAAACGTTAGCCGA TT-3‘. 2.2.19.2: Agarose gel electrophoresis Agarose gel electrophoresis was carried out to analyze the amplified DNA samples. 1.5 percent w/v agarose gel was prepared by melting 1.5 gm of agarose in 100 ml of 1 X TBE buffer in microwave oven. 4 µl (4 µl/ 100 ml) gel red solution was added to stain DNA. DNA samples were mixed with DNA loading dye containing 0.25% bromophenol blue prepared in 40% sucrose solution. Electrophoresis was performed at 100 volts (50 mA) for 65 minutes in 1 X TBE running buffer. After electrophoresis amplified product was detected by placing the gel on UV-Trans illuminator (Life Technology, USA). COMPOSITION OF 10 X TBE Tris 890 mM Boric acid 25 mM EDTA 0.1 mM (pH 8.3) 2.2.20: Southern blot analysis Southern blot analysis of transformants was performed by extracting the genomic DNA from the plant leaves by using CTAB (Cetyl trimethyl ammonium bromide) method of Doyle and Doyle, (1990). The DNA was digested and agarose gel electrophoresis was carried out and the separated DNA fragments were transferred to a positively charged nylon membrane. The probe was prepared and hybridization of membrane carrying plant DNA was done with the probe. Finally the membrane was placed over an x-ray film for exposure. 2.2.20.1: DNA restriction Digestion restriction was carried out by using restriction endonuclease under conditions recommended by the manufacturer (Promega). Approximately 50 µg of genomic DNA isolated from both transformed and untransformed plants were digested with EcoRI in 10 μl reaction: DNA 50 µg KPNI (1 U/μl) 0.2 μl 10X Buffer 1 μl PCR H2O to 10 μl The digestion mixture was incubated at 37 °C for 2 hours in 10 μl reaction. The DNA was subsequently precipitated with ethanol and sodium acetate, resuspended in 15 μl TE buffer after washing with 70% ethanol. 2.2.20.2: Agarose gel electrophoresis DNA loading buffer (5 μl) was added to each sample and separation of digested DNA was carried on 0.8% (w/v) TBE agarose gel containing 0.5 μg/ml (w/v) ethidium bromide. Electrophoresis was carried out for 16 hours at 40mA constant current. 2.2.20.3: Transfer of restriction fragments to membrane The agarose gel containing the separated DNA fragments was treated for 15 min. in 0.25 M HCl at RT, and then shaken for 30 min. in denaturing solution (Appendix-V) and for another 30 min. in neutralizing solution (Appendix-V). The DNA fragments were transferred from the gel to the Hybond-N+ nitrocellulose membrane overnight with 20 X SSC buffer. For fixation of the DNA fragments, the membrane was exposed to UV light for 5 min. and baked at 80°C for 2 h. 2.2.20.4: Labeling of DNA PCR products of rol A gene from plasmid was used as the probe. The probe was labeled using digoxigenin (DIG)-11-dUTP with DIG High Prime DNA Labeling reagents (Roche, Mannheim, Germany). 15 μl probe was added in 25μl sterile distilled water and 10 μl hexanucleotide in a total volume of 50 μl and incubated for 5 min. at 95°C for denaturation. Then it was quickly chilled on ice. Finally, 3 μl mix i.e. 2 μl (DIG)-11-dUTP and 1 μl Klenow enzyme (5 U/μl), were added and incubated for 10 min. at 37°C. After incubation, 4 μl dNTPs was added and it was reincubated for 5 min. at 37°C. The reaction was stopped with 50 μl TE buffer; pH 8 and the probe was allowed to pass through a sephadex column (to clean the probe). Before using the probe, it was incubated at 95°C for 5 min. for denaturation and quickly chilled on ice. 2.2.20.5: Hybridization process The pre-hybridization was performed at 45°C in 50-100 ml hybridization solution (Appendix-V) without adding labeled DNA. After 3 hours the hybridization solution was discarded and replaced with the fresh solution, after adding labeled DNA probe, hybridization was performed overnight at 45°C. Washing of the membrane was carried out thrice with 50 ml washing buffer (20X SSPE, 10% SDS and dH2O 408ml) at 65°C for 20 minutes. Finally, Kodak hyper-film (X-ray) was exposed with the hybridized membrane for 3-5 days at -70°C. 2.3: Results The aim of the present study was to optimize the transformation conditions for A.annua and A.dubia, establishment of hairy root cultures and production of different transgenic lines of both species. Molecular analysis was performed for confirmation of transformation through PCR and Southern blot analysis. Agrobacterium tumefaciens strain LBA4404 containing pRT99 harboring rol genes and Agrobacterium rhizogenes strain LBA8196 and 9402 were used for the transformation experiments. Numbers of factors have been reported that affect the efficiency of transformation with A.tumefaciens and A.rhizogenes, different parameters like seed surface sterilization with different time durations, effect of type of explants and co-cultivation period were studied. 2.3.1: Agrobacterium tumefaciens mediated transformation of Artemisia dubia and Artemisia annua with rol ABC Genes Agrobacterium tumefacienes strain LBA4404 containing pRT99 harboring rol ABC genes were used for the transformation experiments. The A. tumefacienes strain LBA4404 grown overnight in MYA medium. Explants were infected with the bacterial strains following the transformation method described in material and methods. 2.3.2: Seed surface sterilization Seeds of both Artemisia species were surface sterilized with Mercuric chloride (HgCl2) 0.1% (w/v) followed by three times washing with sterilized distilled water. Seeds were exposed to sterilizing agents for varying durations. No contamination was observed in seeds of all treatments. The number of seeds germinated varied with different duration of exposure to the sterilizing agent i.e. Mercuric chloride (HgCl2) 0.1% (w/v). The germinated seeds are shown in figure 2.2 and 2.3. 2.3.3: Medium for seed germination Seeds sown on half strength MS medium showed 100% germination within 3 days. Whereas on plain agar although 80% of the seeds germinated but the duration was almost 6 days (as shown in the table 2.1). Moreover, the seeds germinated on half MS medium were healthier and fresh as compared to the ones on plain agar. Table 2.1: Effect of medium on germination of seeds. Medium Duration of germination Percentage of seeds germinating ½ MS 3 days 100% Plain agar 6 days 80% 120 Percentage of seed germination % age seed germination 100 80 60 40 20 0 2 min 5 min 10 min Sterilizartion time 20 min Fig 2.2: Percentage seed germination with different duration of exposure to 0.1% (w/v) Mercuric Chloride (a) (c) (b) (d) Fig 2.3: In vitro grown seedlings of Artemisia dubia (a, b) and Artemisia annua (c, d) on MS. 2.3.4: Types of explants The Agrobacterium tumefaciens mediated transformation procedure involving tissue culturing was carried out with two different types of explants i.e. stem and leaf with one strain of Agrobacterium tumefaciens i.e. LBA4404 containing pRT99 harboring rol ABC genes. Explants were prepared by cutting leaf and stem segments from 20 days old seedlings (as shown in figure 2.4) into 0.5-1 cm pieces and transferred to MS medium containing 200µM acetosyringone. These pieces were used as explants for further manipulation. (a) (b) (c) (d) Fig 2.4: Leaf and stem explants used for transformation Artemisia dubia (a, b) and Artemisia annua (c, d) 2.3.5: TRANSFORMATION 2.3.5.1: Effect of co-cultivation period on transformation Explants from 20 days old in vitro grown seedlings (Fig. 2.4) were pre-cultured for three days on co-cultivation medium containing MS with 200µM acetosyringone. After pre-culturing, explants were co-cultivated with bacterial culture and were put on the co-cultivation medium. After two days on co-cultivation medium, explants were shifted to the selection regeneration media. Leaf and stem explants were incubated for 3, 5, 10 and 15 minutes in Agrobacterium solution as shown in fig 2.5 a (A.dubia) and 2.5 b (A.annua). Highest (70%) regeneration response was observed in explants incubated for 5 minutes. Further increase in the length of incubation decreased the transformation frequency as shown in figure 2.6. 2.3.5.2: REGENERATION 2.3.5.2.1: Effect of antibiotics on Agrobacterium in regeneration medium To control Agrobacterium growth in selective regeneration medium after transformation cefotaxime 500 mg/L was used. Cefotaxime (Claforan®, Hoechst AG, and Frankfurt, Germany) shows good antimicrobial activity against Agrobacteria strains and exhibits an auxin like activity on plant material Vergauwe et al. (1996b). In some experiments Agrobacterium growth occurred again when regenerated transformed shoots were transferred to fresh regeneration medium without cefotaxime after 4 weeks and these regenerated shoots died due to over growth of Agrobacterium. Therefore, in other experiments cefotaxime was used for two months after their transfer to regeneration medium 2.3.5.2.2: Regeneration of transgenic plants Leaf and stem explants excised from one week old seedlings were cocultured with Agrobacterium on cocultivation medium. After three days these explants were washed with sterile distilled water and then with cefotaxime to remove excess of Agrobacterium. Leaf apices were cultured for direct shooting on SRM 2 containing 0.1mg/L BAP supplied with Kanamycin 20mg/L and cefotaxime 500mg/L. After two weeks transgenic shoots were obtained (Fig. 2.7). (a) (c) (b) (d) Fig 2.5 (a): Regeneration response of Artemisia dubia explants on selection medium after incubation in Agrobacterium solution for various durations. (a). 3 min (b). 5 min (c). 10 min (d). 15 min (a) (c) (b) (d) Fig 2.5 (b): Regeneration response of Artemisia annua explants on selection medium after incubation in Agrobacterium solution for various durations. (a). 3 min (b). 5 min (c). 10 min (d). 15 min Fig 2.6: Effect of incubation period in Agrobacterium solution on regeneration efficiency of explants. (a) (b) © (d) Fig. 2.7: Shooting of transformants from different explants Artemisia dubia (a) Leaves (b) Stem Artemisia annua (c) Leaves (d) Stem 2.3.5.3: Rooting of transformed plants For rooting, when the transgenic shoots were elongated to 2-4cm, the developed shoots of Artemisia dubia cut off segments were cultured on rooting medium (RM) i.e. half strength MS (Murashige and Skoog, 1962) medium containing 0.025 mg/L NAA while developed shoots of Artemisia annua were cultured on rooting medium (RM) i.e. full strength MS (Murashige and Skoog, 1962) medium containing 0.1 mg/L NAA. Rooting was quantified on the basis of percentage of shoots showing response for rooting. The rooting response in transformed plants was found to be 50% and has been shown in figure 2.8. (a) (b) © (d) Fig 2.8: Rooting response of transformed plants Artemisia dubia (a, b) and Artemisia annua (c, d) 2.3.6: Morphological analysis of transformed plants Clear morphological differences were found in some of the green house grown transgenic plants as compared to control plants of both Artemisia species. The morphological differences studied had been listed in table 2.2 and represented in figure 2.9. Control plants were healthier and survived longer as compared to transgenic plants. Marked differences exist in height between control and transgenic plants growing in a greenhouse. Table 2.2: Morphological differences observed among Transgenic and Control Plants Morphological Control Plants (cm) + S.E characters Average Transgenic Plants (cm) + S.E Plant 87cm + 0.5 70cm + 0.3 height Stem Straight, unbranched and soft in Branched and hard in texture texture Leaves Large size and broad Small size and narrow Inflorescence Axial, without hairs Terminal, excessive hairy (a) (b) (c) (d) Fig. 2.9: Comparison of transformed and non-transformed plants (a) Non-Transformed plant of A. dubia (b) Transformed plant of A. dubia (c) Non-Transformed plant of A. annua (d) Transformed plant of A. annua Agrobacterium rhizogenes mediated transformation of Artemisia dubia and Artemisia annua with rol genes Agrobacterium rhizogenes strain LBA9402 and LBA8196 rol genes were used for the transformation experiments. The A. rhizogenes strains grown overnight in MYA medium. Plants were infected with the bacterial strains following the transformation method described in material and methods. 2.3.7: In-vitro plant production and sterilization For Agrobacterium rhizogenes mediated transformation, small stem explants carrying a bud were excised from green house grown plants of both Artemisia species and cultured on MS medium. New plants appeared from bud after one month; these plantlets were cut and then transferred to half MS medium (Fig 2.10). The subculturing of these plantlets to half MS medium resulted in the propagation of multiple plantlets from a single bud. (a) (b) Fig. 2.10: In-vitro grown plantlet of (a) Artemisia dubia (b) Artemisia annua 2.3.8: Production of hairy roots The four-week old plants of both Artemisia species from half MS medium were infected with Agrobacterium rhizogenes strains LBA9402 and LBA8196; hairy roots appeared after six days of infection (Fig. 2.11), whereas no hairy roots were produced with the strain LBA8196. 2.3.9: Molecular analysis of transformed plants Molecular analysis of the seedlings transformed by rol ABC genes, obtained from tissue culturing was performed. 2.3.9.1: Genomic DNA extraction For PCR analysis, genomic DNA was extracted using CTAB method (Cetyl trimethyl ammonium bromide) of Doyle and Doyle, (1990) from the leaves of the transformed plants. DNA extracted through this procedure was run through the Gel (1% agarose) and visualized under U.V transilluminator. (a) (b) Hairy Roots © (d) Fig. 2.11: Induction of hairy roots from stems infected with Agrobacterium rhizogenes. (a, b) A. dubia (c, d) A. annua 2.3.9.2: Polymerase chain reaction (PCR) For molecular analysis, genomic DNA was isolated by using CTAB method of Doyle and Doyle, (1990) from transformed and untransformed plants of both Artemisia species, and plasmid DNA was also isolated by using alkaline lysis method from both Agrobacterium tumefaciens and Agrobacterium rhizogenes. PCR was performed for rol A, B and C genes and the amplified products (308bp, 779bp, 540bp respectively) were observed to confirm transformation. Same size amplified product was also obtained from the plasmid DNA of Agrobacterium strains. Rol A, B and C genes were detected in all transformed lines of Artemisia dubia (D1, D2, D3, DH1 and DH2) and Artemisia annua (A1, A2, A3, AH1 and AH2) as well as in plasmid DNA but not in control plants (Fig. 2.12a and 2.12b). The figure 2.12 shows the amplified products of rol A, B and C genes. M P D1 D2 D3 DH1 DH2 C 3000 2000 1000 308bp 750 500 250 (a) M C P D1 D2 D3 DH1 DH2 3000 2000 1000 750 500 250 779bp (b) M C P D1 D2 D3 DH1 DH2 3000 2000 1000 750 540bp 500 (c) 250 Fig. : 2.12 (A). PCR analysis showing amplified product of (a) rol A (b) rol B and (c) rol C in transgenic Artemisia dubia 3000 M C P A1 A2 A3 AH1 AH2 2000 1000 750 500 308bp 250 (a) M C AI A2 A3 AH1 AH2 P 3000 2000 1000 750 779bp 500 250 (b) M C AI A2 A3 AH1 AH2 P 3000 2000 1000 750 540bp 500 (c) 250 Fig. : 2.12 (B). PCR analysis showing amplified product of (a) rol A (b) rol B and (c) rol C in transgenic Artemisia annua Fig 2.12: PCR analysis showing amplified products (A) D1-D3 represents the plants transformed by Agrobacterium tumefaciens and DH1-DH2 represents the plants transformed by Agrobacterium rhizogenes and in Artemisia annua (B) A1-A3 represents the plants transformed by Agrobacterium tumefaciens and AH1-AH2 represents the plants transformed by Agrobacterium rhizogenes. Lane P represents the plasmid DNA. Lane C refers to the non transformed control plants. Lane M corresponds to 1 kbp Ladder (Fermentas). 2.3.10: Southern blotting Southern blot analysis of PCR-positive plants of Artemisia dubia and Artemisia annua was performed by DIG High Prime DNA Labeling and Detection Starter Kit II (Roche Cat. No. 11585614910) according to the manufacturer‘s instructions. Hybridization bands were detected in Southern blots with the rol A probe in the transgenic plants. In all the transformants, the inserted copy number was one except line D2 of Artemisia dubia and A1 and A3 of Artemisia annua in which two copies of the inserted genes were observed (Fig. 2.13 a,b). The results confirmed the integration of the Agrobacterium T-DNA in the genome of Artemisia dubia and Artemisia annua transgenic lines. C D1 D2 D3 DR1 DR2 4.5 kb 4.0 kb 3.5 kb 3.0 kb 2.5 kb 2.0 kb (a) C A1 A2 A3 AH1 AH2 4.5 kb 4.0 kb 3.5 kb 3.0 kb 2.5 kb 2.0 kb (b) Fig. 2.13: Southern blot analysis of (a) Artemisia dubia (b) Artemisia annua. C untransformed control plant, lanes 1–5 rol ABC transformed plants (Artemisia dubia), lanes 1-5 rol ABC transformed plants (Artemisia annua), D1-D3 represents the plants transformed by Agrobacterium tumefaciens and DH1-DH2 represents the plants transformed by Agrobacterium rhizogenes (Artemisia dubia), A1-A3 represents the plants transformed by Agrobacterium tumefaciens and AH1-AH2 represents the plants transformed by Agrobacterium rhizogenes (Artemisia annua). 2.3.11: Acclimatization Transformed rooted plantlets of both Artemisia species were shifted to small pots containing a mixture of soil and peat moss. Plants were then acclimatized in larger pots in growth chamber at 25 ±2ºC, 16 h of photoperiod, illumination of 45 uEm-² s-¹ and 60% and 60 % relative humidity before transfer to the harsher conditions of the green house environment. Out of 15 plants shifted to pots, only 8-10 seedlings were able to survive the growth room conditions, so the percentage of seedlings acclimatized to the outer environment was 53-66%. Rest of the seedlings could not withstand in the growth room (Fig: 2.14a, b). (a) (b) © (d) Fig. 2.14: Transformed plant in green house (a, b) Transformed plant of A. dubia (c, d) Transformed plant of A. annua 2.4: Conclusion In this part, transformation of Artemisia annua and Artemisia dubia with rol genes was carried out through A.tumefacienes and A.rhizogenes and following conclusion were drawn:  Both species of Artemisia (A.annua and A.dubia) have shown variable regeneration responses and both proved best for in vitro regeneration.  Factors like genotype, explants type, cocultivation period and sterilization time affect the in vitro regeneration ability.  Among the two explants (Stem and Leaf) used for in vitro regeneration, maximum shoot formation was observed in stem explants followed by leaf explants.  Maximum shoot regeneration was recorded on MS + 0.1 mg/l BAP + 20mg/ml kanamycin + 500mg/l cefotaxime.  Rooting was observed on half MS medium with 0.025mg/l for A.dubia while half MS medium with 0.1mg/l was used for A.annua shoots.  Agrobacterium mediated transformation was the preferred method of Artemisia transformation dependant on optimum combination of sterilization time, inoculation time and co-cultivation period for stem and leaf explants.  Stable transformation of Artemisia with rol genes was carried out successfully by employing Agrobacterium tumefacienes and Agrobacterium rhizogenes mediated transformation.  Transformation and insertion of desired genes from both transformation methods was confirmed by PCR and southern blotting. Chapter No. 3 Analysis of Artemisinin and its Derivatives and their comparison in transformed and untransformed plants of Artemisia annua and Artemisia dubia 3.1: Introduction Artemisinin an endoperoxide sesquiterpene lactone produced by aerial parts of Artemisia plant is effective even against multi-drug resistant strains of the malarial parasite. The isolation and characterization of artemisinin from Artemisia plant is considered as one of the most novel discoveries in recent medicinal plant research (Roth and Acton, 1989). Although the complete organic synthesis has been established, but chemical synthesis of artemisinin is not yet economically feasible because of its complexity and low yield. Currently the leaves, roots and flowers of Artemisia species form the only source of this drug. Artemisinin is found in very low quantities (0.05%-1.1%) in different cultivars of Artemisia dubia and Artemisia annua (1%-4%). High artemisinin yielding clones are being isolated by selection and other non-conventional approaches; however, these have their own limitations (Delabays et al., 1993). Genetic improvement of natural varieties has been attempted but the maximum yield of artemisinin from this route reached so far is 2% (Graham et al., 2010). Through transformation we can enhance the production of artemisinin in Artemisia dubia and Artemisia annua by inserting our desired genes. A number of publications have reported large differences in artemisinin content depending on variety, season, different plant parts, co cultivation conditions and plant developmental stage (Lommen et al., 2007; Davies et al., 2009; Yang et al., 2009). Artemisinin has been detected in leaves; small green stems, buds, flowers and seeds (Martineg and Staba, 1988; Ferreira et al., 1995, Abdin et al., 2003). Reports on the distribution of artemisinin throughout the plant are however inconsistent. Many analytical procedures to identify and quantify artemisinin, its biosynthetic precursors as well as its metabolites have been developed during the last 30 years. These include thin layer chromatography (Klayman et al., 1984), High performance liquid chromatography with UV detection (Zhao and Zeng, 1985), gas chromatography coupled to mass spectrometry (Woerdenbag et al., 1991), capillary electrophoresis coupled to UV detector (Christen and Veuthey, 2001), HPLC with evaporative light scattering detection (Kohler et al., 1997; Christen and Veuthey, 2001) and enzymelinked immunosorbent assay (Jaziri et al., 1993). The aim of the work presented in this chapter was to analyze and compare artemisinin and its derivatives (artemether, arteether, dihydroartemisinin and artesunate) in different tissues (transformed shoots, transformed roots and hairy roots) of transformed and untransformed plants of Artemisia annua and Artemisia dubia. 3.2: Materials and Methods 3.2.1: Extraction of artemisinin and derivatives of artemisinin Artemisinin and derivatives of artemisinin were extracted from transformed and non-transformed shoots, roots and hairy roots of Artemisia dubia and Artemisia annua by using following method: Biomass extraction was performed according to the method by Lapkin et al. (2006). Basically, 1 g of plant leaves was extracted with 30ml hexane/ethyl acetate (9.5:0.5) at room temperature in a Sonication bath (Kerry Ultrasonic UK) for 1 hour. Hexane and ethyl acetate mixture was used to increase the solubility of artemisinin in the extraction solvent. The extract was filtered using Whatman 1.0um filter paper and concentrated in vaco to give the crude extract (Solvent 1). For quantitative analysis of artemisinin and its derivatives, an aliquoted volume of the crude extracts were dried in a speed vacuum (Jousan RC1022) at room temperature and re-suspended in acetonitrile to dissolve artemisinin and other acetonitrile soluble components of the extract while precipitating waxes and other non-soluble components (Solvent 2). The suspension was filtered with a syringe filter (0.45um Millipore) into an HPLC vial ready for analysis. 3.2.2: Reagents used HPLC grade hexane, ethyl acetate and acetonitrile were obtained from Fisher scientific UK. Artemisinin, artemether, arteether, artesunate and dihydroartemisinin reference standards (>98.0 %) were obtained from Sigma-Aldrich (UK). Deionised water was obtained from Millipore Q-POD (UK) purification unit. 3.2.2.1: Solvents used 3.2.2.1.1: Solvent No.1: Mixture of ethyl acetate and hexane in the ratio of 5:95  For 10 ml we have to add 0.5ml Ethyl acetate and 9.5ml Hexane  For 500 ml we have to add 25 ml Ethyl acetate and 475 ml Hexane 3.2.2.1.2: Solvent No.2: Mixture of acetonitrile and water in the ratio of 50:50  For 10 ml we have to add 5ml Acetonitrile and 5ml Water  For 500 ml we have to add 250 ml Acetonitrile and 250 ml Water Also we can use 100% Acetonitrile 3.2.2.1.3: Biomass and solvent ratio  Biomass and solvent ratio should be 1:10  For 1g plant material we should add 10ml of solvent 3.2.3: Analysis of artemisinin content Artemisinin and its derivatives were extracted from shoots, roots and hairy roots of transformed and untransformed plants of Artemisia dubia and Artemisia annua and HPLC analysis of extracts was performed on a Shimadzu Prominence HPLC-UV/ELSD instrument equipped with auto-sampler, degasser and a photo-diode array detector. The protocol employed was adapted after Lapkin et al. (2006). Separation was achieved on a Zorbax Eclipse XDB column (15 x 4.6cm, 0.5m) using a 50% acetonitrile mobile phase flowing at 1ml/min and Injection volume of 20 µl. Detection and quantification of artemisinin was carried out at 210nm using calibrated standard. 3.2.3.1: High performance liquid chromatography Chromatographic analysis of plant extracts was performed on a Shimadzu Prominence HPLC 20A, equipped with auto-sampler, degasser, photodiode array and evaporative light scattering (ELSD) detector. The protocol employed was modified after Lapkin et al. (2009). The separation and identification of artemisinin and its derivatives in extracts were achieved by a 50:50 acetonitrile/water mobile phase flowing at 1ml/min on an Agilent Eclipse Zorbax column (15 x4.6cm, 0.5um). Due to absorption at low UV wavelength of artemisinin and its derivatives, ELSD was used in series with UV for the detection and quantification. Using this protocol, the artemisinin, artemether, arteether, artesunate and dihydroartemisinin peaks were well resolved therefore there was no need for a buffered mobile phase to aid separation. Identification of peaks in extracts was analyzed by comparison with retention indices of reference standards and by spiking of extracts with appropriate standards. Before the injection of samples and standards the HPLC system was allowed to pump the mobile phase and to equilibrate for 1hour until a flat consistent baseline was achieved. For safe use of HPLC system, every day, after last sample run, mobile phase was run again followed by run of 50 % and 100 % methanol for 30 minutes each. The mobile phase used was at pH 7.0, because a higher pH can dissolve the column. 3.2.3.1.1: Preparation of mobile phase Mobile phase was prepared by adding together 50:50 acetonitrile/water and filtered with 0.22 μm aqueous filter (cellulose acetate membrane, Sartorius). Furthermore, mobile phase was sonicated before use for 40-45 minutes to remove the air that may interrupt the normal operations of HPLC as well as results of samples. 3.2.3.1.2: Preparation of dilutions of standard 3.5mg/ml of artemisinin, artemether, arteether, artesunate and dihydroartemisinin reference standards were prepared in acetonitrile from which a serial dilution of the standards were made to give the following concentrations of 1.75, 0.70, 0.35, 0.18 and 0.04 mg/ml (or 12.21, 6.21, 2.48, 1.24, 0.62 and 0.12 M) of artemisinin. HPLC analysis of the dilutions produced a calibration graph from which the qualitative and quantitative analysis of artemisinin and its derivatives in crude extracts were determined. 3.2.3.1.3: Column used for HPLC Samples were analyzed by HPLC using Eclipse Zorbax SB C18 column (150 x 4.6mm x 5μm) of Agilent technologies. 3.2.3.1.4: Flow of mobile phase through HPLC system Flow of mobile phase through column (stationary phase) was 1 ml/min. 3.2.3.1.5: Detector Diode Array Detector (G1315B-DAD) showed maximum absorbance of artemisinin and its derivatives at 260 nm. 3.2.3.1.6: Retention time of peak Retention / elution time peak of artemisinin and derivatives was 40 minute. Artemisinin and its derivatives were identified by using the authentic standard of artemisinin and its derivatives. 3.2.3.1.7: Injection volume An injection volume of 20 μl of sample as well as standard was used. 3.2.3.1.8: Calibration curve Six dilutions1.75, 0.70, 0.35, 0.18 and 0.04 mg/ml (or 12.21, 6.21, 2.48, 1.24, 0.62 and 0.12 M) of standard artemisinin, artemether, arteether, artesunate and dihydroartemisinin were used. The calibration curves were obtained by plotting the chromatographic peak area (mAU) against the concentration (g/ml), where a linear response was observed. Linear regression analysis resulted in the equation with an excellent correlation coefficient i.e. R2 = 1. The calibration curve and linear regression equation were used for the determination of artemisinin in the samples (Guo et al., 2005). The linear regression equation according to standard concentrations (g/mL) is as under: y = mx + b m = 1.69159 e–1 b = 7.20139e–1 y = 1.69159 e–1x + 7.20139e–1 Where; y = values [Peak Area (mAU)] on y-axis x = values [Concentration (g/ml)] on x-axis 3.2.3.1.9: Quantification of artemisinin and its derivatives in sample For any given Peak Area (mAU) i.e. y from sample chromatogram, corresponding concentration (g/mL) of sample can be calculated by evaluating x as follows: x = (y - 7.20139e–1) / 1.69159 e–1. 3.2.3.1.10: Verification of artemisinin and derivatives of artemisinin concentrations Co-injection was used in order to positively identify peaks as artemisinin, artemether, arteether, artesunate and dihydroartemisinin. Suspected artemisinin and its derivatives peaks were noted, and their area recorded from the original sample that was run. An aliquot of a concentrated, Q260 derivatized, filtered (0.22 µm FP-200 13mm FPVericel™ membrane filter, Pall-Gelman Laboratory) artemisinin, artemether, arteether, artesunate and dihydroartemisinin standard (Sigma-Aldrich, St. Louis, MO) solutions were added to the analyzed sample to achieve an artemisinin and derivatives peak area about four times that of the area of the putative artemisinin and derivatives peak. If the putative peak increased in area to an amount that was 4-5 times its original area, the peak was determined to be that peak respective to its standard. Quantitation was based on the original injection prior to the addition of the standard. 3.2.4: Statistical analyses All experiments were done at least in triplicate for each tissue, using an average artemisinin concentration of three plants of each species. Data were statistically analyzed using ANOVA and Duncan‘s Multiple Range Test. 3.3: Results 3.3.1: Analysis of artemisinin content in transformed and untransformed Artemisia annua and Artemisia dubia Artemisinin was analyzed in plants of Artemisia dubia and Artemisia annua. Artemisinin was extracted from shoots, roots and hairy roots by using method described in materials and methods (Section: 3.2.1). Artemisinin was detected on HPLC at 210nm and retention time of 10.17 in case of A.annua plants (Fig. 3.1) and 9.24 in case of Artemisia dubia (Fig 3.2). The average artemisinin content from three transgenic lines of Artemisia annua varied between tissues, the difference between the lines is shown in fig. 3.3. In the shoots this was 46.89 mg/g dry weight (DW) compared to 4.37 mg/g DW in the controls shoots. Roots of transformed plants had an average artemisinin content of 3.06 mg/g DW while the amount found in transformed hairy roots was significantly greater (15.37 mg/g DW) compared to the control roots in which negligible amounts could be detected (0.02 mg/g DW). The average artemisinin in transgenic lines of A. dubia was lower than A. annua this was 3.78 mg/g DW in shoots compared to 0.11 mg/g DW in control shoots. The transformed plant roots showed 1.17 mg/g DW and hairy roots 1.62 mg/g DW of A.dubia, whereas no artemisinin was found in control roots of A. dubia (Fig. 3.3). In general, higher amount of artemisinin was observed in transgenic plants of Artemisia annua as compared to transgenic plants of Artemisia dubia (Fig. 3.3). (a) (b) (c) Fig. 3.1: HPLC Chromatograms of (a) Standard Artemisinin (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Artemisinin) (a) (b) (c) Fig. 3.2: HPLC Chromatograms of (a) Standard Artemisinin (b) Artemisia dubia Sample (D2) (c) Spike Sample (A.dubia sample + standard Artemisinin) 60 Artemisinin content (mg/g) DW 50 40 30 20 10 0 Fig. 3.3: Comparative analysis of Artemisinin content in shoots, roots and hairy roots of transformed and un-transformed plants of Artemisia dubia and Artemisia annua. The average of three plants is shown together with error bars showing SE. A2+Artemisinin and D2+ Artemisinin are the spike samples used for verification of artemisinin concentration 3.3.2: Analysis of derivatives of artemisinin (Artemether, Arteether, Artesunate and Dihydro-artemisinin) in transformed and untransformed Artemisia annua and Artemisia dubia plants Artemether, arteether, artesunate and dihydro-artemisinin were analyzed from plants of Artemisia dubia and Artemisia annua. These were extracted from shoots, roots and hairy roots by using method described in materials and methods (Section: 2.23). Artemether, arteether, dihydroartemisinin and artesunate were detected on HPLC at 210nm and retention time of 7.57, 2.85, 4.79, and 1.94 respectively in case of A.annua plants (Fig. 3.4, 3.6, 3.8, 3.10) and 7.51, 2.88, 4.97, 1.84 respectively in case of Artemisia dubia (Fig. 3.5, 3.7, 3.9, 3.11). The average artemether content from three transgenic lines of Artemisia annua varied between tissues, the difference between the lines is shown fig. 3.12. In the shoots this was 4.3 mg/g dry weight (DW) compared to 1.05 mg/g DW in the controls shoots. Roots of transformed plants had an average artemether content of 1.5 mg/g DW while the amount found in transformed hairy roots was significantly greater (4 mg/g DW) compared to the control roots in which negligible amounts could be detected (0.01 mg/g DW). The average artemether in transgenic lines of A. dubia was lower than A. annua this was 2 mg/g DW in shoots compared to control shoots (0.3 mg/g DW), transformed plant roots (0.7 mg/g DW) and hairy roots (1.62 mg/g DW), whereas negligible amount of artemether was found in control roots of A. dubia (Fig. 3.12). In general, higher amount of artemether was observed in transgenic plants of Artemisia annua as compared to transgenic plants of Artemisia dubia (Fig. 3.16). The average arteether content from three transgenic lines in the shoots of Artemisia annua was 7 mg/g dry weight (DW) compared to 1.5mg/g DW in the controls shoots. Roots of transformed plants had an average arteether content of 1.6 mg/g DW while the amount found in transformed hairy roots was significantly greater (3.6 mg/g DW) compared to the control roots in which negligible amounts could be detected (0.016 mg/g DW). The average arteether in transgenic lines of A. dubia was lower than A. annua this was 3.16 mg/g DW in shoots compared to control shoots (0.1mg/g DW), transformed plant roots (0.75 mg/g DW) and hairy roots (1.61 mg/g DW), whereas negligible amount of arteether was found in control roots of A. dubia (Fig. 3.13). In general, higher amount of arteether was observed in transgenic plants of Artemisia annua as compared to transgenic plants of Artemisia dubia (Fig. 3.16). The average Dihydro-artemisinin content from three transgenic lines in the shoots of Artemisia annua was 8.9 mg/g dry weight (DW) compared to 2.2 mg/g DW in the controls shoots. Roots of transformed plants had an average dihydro-artemisinin content of 2.5 mg/g DW while the amount found in transformed hairy roots was significantly greater (6.8 mg/g DW) compared to the control roots in which negligible amounts could be detected (0.03 mg/g DW). The average artemisitene in transgenic lines of A. dubia was lower than A. annua this was 3.5 mg/g DW in shoots compared to control shoots (0.25 mg/g DW), transformed plant roots (0.91 mg/g DW) and hairy roots (1.53 mg/g DW), whereas negligible amount of dihydro-artemisinin was found in control roots of A. dubia (Fig. 3.14). In general, higher amount of dihydroartemisinin was observed in transgenic plants of Artemisia annua as compared to transgenic plants of Artemisia dubia (Fig. 3.16). The average artesunate content from three transgenic lines in the shoots of Artemisia annua was 5.5 mg/g dry weight (DW) compared to 1.04 mg/g DW in the controls shoots. Roots of transformed plants had an average artesunate content of 0.86 mg/g DW while the amount found in transformed hairy roots was significantly greater (2.6 mg/g DW) compared to the control roots in which negligible amounts could be detected (0.006 mg/g DW). The average artesunate in transgenic lines of A. dubia was lower than A. annua this was 2.16 mg/g DW in shoots compared to control shoots (0.10 mg/g DW), transformed plant roots (0.61 mg/g DW) and hairy roots (1.44 mg/g DW), whereas negligible amount of artesunate was found in control roots of A. dubia (Fig. 3.15). In general, higher amount of artesunate was observed in transgenic plants of Artemisia annua as compared to transgenic plants of Artemisia dubia (Fig. 3.16). (a) (b) (C) Fig. 3.4: HPLC Chromatograms of (a) Standard Artemether (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Artemether) (a) (b) © Fig. 3.5: HPLC Chromatograms of (a) Standard Artemether (b) Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Artemether) (a) (b) (C) Fig. 3.6: HPLC Chromatograms of (a) Standard Arteether (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Arteether) (a) (b) (C) Fig. 3.7: HPLC Chromatograms of (a) Standard Arteether (b) Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Arteether) (a) (b) (C) Fig. 3.8: HPLC Chromatograms of (a) Standard Dihydroartemisinin (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Dihydroartemisinin) (a) (b) (C) Fig. 3.9: HPLC Chromatograms of (a) Standard Dihydroartemisinin (b) Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Dihydroartemisinin) (a) (b) (C) Fig. 3.10: HPLC Chromatograms of (a) Standard Artesunate (b) Artemisia annua Sample (A2) (c) Spike Sample (A.annua sample + standard Artesunate) (a) (b) (C) Fig. 3.11: HPLC Chromatograms of (a) Standard Artesunate (b) Artemisia dubia Sample (D1) (c) Spike Sample (A.dubia sample + standard Artesunate) 8 Artesunate Content (mg/g) DW 7 6 5 4 3 2 1 0 Fig. 3.12: Artemether content in shoots roots and hairy roots of A.annua.A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Artemether and D2+ Artemether are the spike samples used for verification of artemether concentration. 10 Arteether content (mg/g) DW 9 8 7 6 5 4 3 2 1 0 Fig. 3.13: Arteether content in shoots roots and hairy roots of A.annua. A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Arteether and D2+ Arteether are the spike samples used for verification of arteether concentration. 12 Dihydro-artemisinin content (mg/g) DW 10 8 6 4 2 0 Fig. 3.14: Dihydro-artemisinin content in shoots roots and hairy roots of A.annua.A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Dihydroartemisinin and D2+ Dihydroartemisinin are the spike samples used for verification of dihydroartemisinin concentration. 8 Artesunate Content (mg/g) DW 7 6 5 4 3 2 1 0 Fig. 3.15: Artesunate content in shoots roots and hairy roots of A.annua.A1, A2 and A3 represents leaf samples from three different transgenic lines, AR1, AR2 and AR3 represents roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots. A2+Arteether and D2+ Arteether are the spike samples used for verification of arteether concentration. Contents of Artemisinin and its Derivatives (mg/g) DW 50 A 45 Artemisinin Artemether Dihydro-artemisinin Artesunate Arteether 40 35 30 25 20 B 15 10 5 C CD EF EF FGHI FGHIJGHIJKL FGHIJKL KLHIJKL EFGH EFGH M HIJKL D EFGH EFGFGHI FGHIJ FGHI GHIJKLFGHIJK FGHIJK N N IJKLKLKLJKLL 0 Fig. 3.16: Comparative and Statistical analysis of Average Artemisinin, Artemether, Arteether, Dihydro-artemisinin and Artesunate content in shoots, roots and hairy roots of transformed and un-transformed plants of A.dubia and A.annua. Each value is the mean of three replicates. Any two means having a common alphabet are not significantly different at p = 0.05 using LSD. Vertical bar represents the standard error of the 3 means. IJKL HIJKL IJKL IJKL JKLM 3.3.3: Statistical Analysis 3.3.3.1: Comparative analysis of artemisinin and its derivatives in different tissues of Artemisia annua and Artemisia dubia Statistical analysis was conducted in factorial design (2 X 5 X 3) to see the effect of different tissues on enhanced production of artemisinin and its derivatives in different tissues of A.annua and A.dubia. Artemisia annua and Artemisia dubia had a significant difference (P<0.05) in production of artemisinin and its derivatives. Most effective results were obtained in case of different tissues of A.annua species (Table 3.1). Table 3.1: Analysis of Variance Table for factors affecting Production of Artemisinin and its Derivatives in different tissues of A.annua and A.dubia Source of Variation df Sum of Squares Mean Square F-Value Prob. Species 1 743.849 743.849 445.5682 0.0000 4 1225.680 306.420 183.5467 0.0000 4 1098.772 274.693 164.5421 0.0000 Tissues 2 881.418 440.709 263.9862 0.0000 Species X Tissues 2 407.269 203.635 121.9780 0.0000 8 1012.384 126.548 75.8028 0.0000 8 970.030 121.254 72.6314 0.0000 Error 60 100.166 1.669 Total 89 6439.567 Artemisinin and its Derivatives Species X Artemisinin and its Derivatives Tissue X Artemisinin and its Derivatives Species X Tissues X Artemisinin and its Derivatives Coefficient of Variation: 27.35% Rol genes had significant impact (P<0.05) on production of artemisinin and its derivatives. Mean value of artemisinin was higher in all plants tissues as compared to other derivatives of artemisinin. Generally artemisinin content was more enhanced in all tissues of both Artemisia species. Similarly, different tissues had significant impact (P<0.05) on production of artemisinin and its derivatives in A.annua and A.dubia. Mean value of artemisinin and its derivatives was higher in shoots of both species as compared to other tissues (Table 3.1). In general Artemisia annua is the best species for production of artemisinin and its derivatives, and among all metabolites artemisinin was most enhanced metabolite compared to other metabolites under the effect of transformation with rol genes. Shoots are the best tissues where production of these metabolites was more enhanced compared to other tissues. Among roots, hairy roots of both species produced more artemisinin and its derivatives compared to other roots. 3.4: Conclusion In this part, Comparative analysis of artemisinin content and its derivatives in transformed and untransformed plants of Artemisia annua and Artemisia dubia was carried out and following conclusion were drawn:  Artemisinin was significantly increased in the transformed shoots, hairy roots and roots of transformed shoots compared to control in both A.annua and A.dubia plants.  Highest amount of artemisinin was found in transgenic shoots of transformed A.annua and A.dubia (increased by a factor of 10) compared to control shoots of both species. Roots of these transformed shoots also contain significantly increased amount of artemisinin compared to control roots.  Significant amount of artemisinin was observed in hairy roots of both A.annua and A.dubia compared to control roots.  Artemether, arteether, dihydroartemisinin and artesunate contents were also significantly increased in transgenic shoots of both Artemisia species compared to control shoots. Amount found in roots of transformed shoots and transformed hairy roots was significantly greater compared to the control roots in which negligible amounts was detected.  In general, results showed that artemisinin content was enhanced much more as compared to its derivatives under the effect of transformation with rol genes and among two Artemisia species, A.annua is the best species for the production of artemisinin and its derivatives. Considering the tissue types, shoots and hairy roots are more appropriate for production of artemisinin and its derivatives. Chapter No. 4 Analysis of Metabolic pathway and Trichome development and their comparison in transformed and un-transformed plants of Artemisia annua and Artemisia dubia 4.1: Introduction Artemisinin is a secondary metabolite that has been found to have strong antimalarial activity with little or no side effects (Klayman, 1985). Artemisinin is a sesquiterpene lactone which has been reported to be produced within the glandular trichomes of Artemisia species, a member of the Asteraceae family. The current understanding is that artemisinin is produced in 10-celled glandular trichomes located on leaves, floral buds, and flowers (Ferreira et al., 1995; Tellez et al., 1999; Olsson et al., 2009) and sequestered in the epicuticular sac at the apex of the trichome (Olsson et al., 2009). The evidence in its support is the observation that, artemisinin concentrations are higher in leaves that are formed later in development than in leaves formed early in the plant‘s development; this difference has been attributed to a higher trichome density and a higher capacity per trichome in the upper leaves (Lommen et al., 2006). It has been shown that hairy roots produced by infection with A.rhizogenes can produce artemisinin (Jaziri et al., 1995; Liu et al., 1999). This would suggest that the plant may be capable of producing artemisinin in the absence of trichomes. There are two independent and differentially-localized pathways involved in terpene biosynthesis that converge to yield a common pool of the terpenoids precursor called isopentenyl diphosphate (IPP) (Croteau et al., 2000; Lange et al., 2000) (Figure 1.15). The mevalonic acid pathway (MVA) which initiates from acetyl-CoA is located in the cytosol. The key, regulatory, step in this pathway is the conversion of hydroxymethylglutaryl-CoA (HMG) to mevalonate via, a regulatory enzyme, hydroxymethylglutaryl-CoA reductase (HMGR). Several subsequent steps lead to formation of the cytosolic localized pool of IPP (Towler and Weathers, 2007). The other pathway, non-mevalonate pathway (MEP), to IPP begins with pyruvate and occurs in the plastid with no mevalonate intermediate. Several subsequent steps synthesize the plastid pool of IPP (Rohmer et al., 1996). Numerous investigations have shown that the Agrobacterium rol genes can induce high levels of secondary metabolites in hairy root cultures of most transformed plant species (Giri and Narasu, 2000; Sevon and Oksman-Caldentey, 2002). We hypothesized that transformation with the rol genes could induce an increase in artemisinin through stimulation of the synthesis pathways. Hence, the aim of the investigation reported here was to compare trichome densities in transformed and untransformed plants of Artemisia annua and Artemisia dubia and the genes involved in the synthesis pathway were monitored to ascertain how the rol genes could act to influence the production of artemisinin. 4.2: Materials and Methods 4.2.1: Analysis of metabolic pathway Young leaves, roots and hairy roots were collected separately from 3-4 months old plants, frozen in liquid nitrogen, ground to a fine powder with electrical homogenizer and used for RNA extraction. 4.2.1.1: RNA extraction RNA extraction was performed using RNAqueous Small Scale Phenol-Free Total RNA Isolation Kit (Ambion) for small scale RNA isolation according to the manufacturer‘s instructions. Frozen plant tissue powder (70 mg) was used for each RNA extraction.  Before working with RNA, lab bench, and pipettes were cleaned with an RNAase decontamination solution such as Ambion RNAase Zap RNAase decontamination solution  70 mg of frozen leaves, roots and hairy roots were weighed and stored at 20ºC for two or three days before extraction.  Elution solution (typically 50-200µl per prep) was pre-heated in an RNAase free microcentrifuge tube in the water bath at 70-80 ºC before starting the extraction.  Samples were ground with freeze electrical grinder, and Lysis / Binding solution in the ratio of 50mg: 600µl was added after grinding.  After adding the solution, samples were centrifuged for 10.minutes at 13,000rpm to remove debris.  After centrifugation, supernatant was transferred to new tube and an equal volume of 64% Ethanol was added and mixed by vortexing.  (a) Lysate /Ethanol mixture was drawn through a filter cartridge. (b) The Maximum volume that can be applied at one time is 700µl; it was centrifuged for 30.second-1 minute at 13000rpm until the lysate/ethanol mixture was through the filter (c) Flow through was discarded and collection tube was reused for the washing steps  700µl Wash solution # 1 was added to the filter cartridge and centrifuged for 30.second, flow through was discarded and tube was reused for subsequent washes.  500µl Wash solution # 2/3 was added and centrifuged for 30.second. Filter cartridge was placed into new collection tube.  Pre-heated elution solution was added to the filter cartridge in two steps, (i) first by adding 40 µl and then centrifuged for 1.minute (ii) second with 30 µl and again centrifuged for 1 minute.  Quality and quantity of RNA was checked at 260 nm by Nanodrop and samples were stored at -20ºC in refrigerator.  Quality of RNA was analyzed by performing RNA 6000 Nano Assay by using Agilent RNA 6000 Nano Assay Kit and Bioanalyzer (Agilent Technologies, Germany, reorder number 5067-1511) according to manufacturer‘s instructions. 4.2.1.2: cDNA synthesis RNA (1 µg) was reverse transcribed using ThermoScript RT-PCR system (Invitrogen) primed with oligo (dT) 20 primer. The RNA was removed from the first strand cDNA by RNase treatment using RNase H (Invitrogen) according to the manufacturer‘s instructions.  cDNA was prepared from RNA by the following method: Table 4.1: Different components used for Sample preparation Amount used for Amount used for 1 reaction 12 reactions S.No. Components 1 DEPC Water 8µl 96µl 2 5mMPrimer (Oligo dT) 1µl 12µl 3 10mM RNA sample 1µl 12µl 4 10mM dNTPs mix 2µl 24µl  All the mixture (prepared above) was denatured by incubating at 65ºC for 5 minutes in pre-heated water bath at 65 ºC and placed on ice after 5 minutes.  Then reaction mixture was prepared for cDNA Synthesis, 5X cDNA Buffer was vortexed for 5 second just prior to use.  Master Mix was prepared and all reactions were performed on ice during this synthesis.  Then the reaction mixture was made as: Table 4.2: Different components used for Master Mix for cDNA Synthesis Amount used for 1 Amount used for 12 reaction reactions 5x cDNA Synthesis Buffer 4µl 48µl 2 0.1 M DTT 1µl 12µl 3 RNAase OUT 1µl 12µl 4 DEPC Water 1µl 12µl 5 Thermo script RT 1µl 12 µl S.No. Components 1  8 µl of reaction mixture was added into the each tube containing sample.  Samples were transferred to the preheated thermo cycler to the appropriate cDNA Synthesis temperature and incubated as follows: (1) 60 minutes at 50ºC (2) 85ºC for 5 minutes.  1 µl of RNAase H was added to the samples and incubated at 37 ºC for 20.minutes.  Samples were stored at -20 ºC immediately after preparation of cDNA or used for PCR immediately. 4.2.2: Quantitative real time polymerase chain reaction (qRT-PCR) The real-time RT-PCR analysis was performed in a LightCycler® 480 RealTime PCR System, version 1.5 (Roche Applied Sciences, Mannheim, Germany), using the cDNAs as templates and the random primers for the analyzed gene, as listed in Table 4.3. The actin gene was used as a reference gene. SYBR Green LightCycler® 480 DNA SYBR Green I Master (Roche Applied Sciences, Mannheim, Germany, Cat. No. 04707516001) was used in the Polymerase Chain Reaction (PCR) to quantify the amount of dsDNA. PCR was performed using 2.5 – 3µl of cDNA in a total of 20µl reaction volume. The PCR conditions were 2 minutes at 95ºC, 30 seconds at 95ºC, 30 seconds at 60-62ºC, 1 minute at 72ºC for 40 cycles, followed by 5 minutes at 72ºC. The conditions were selected because none of the samples analyzed reached a plateau at the end of the amplification (i.e. they were at the exponential phase of the amplification). Table 4.3: Primers used for qRT-PCR No Name 1 TFAR1 2 ADS 3 ALDH1 4 CYP71AV1 5 ACTIN Forward Primer Reverse Primer CCTTGGAGATCCTGAA CGTTGGATTGTGCTGAAC GCTG TG GGGAGATCAGTTTCTC CTTTTAGTAGTTGCCGCA ATCTATGAA CTTCTT CAGGAGCTAATGGAA TTTCTTCCTTCGGCCACTG GTTCTAAGTCAG TTG AGGGTAGGCATTCGCC TCGAGTGGCCCTAACAAC GTCC CTGC ATCAGCAATACCAGG AGGTGCCCTGAGGTCTTG GAACATAGT TTCC Accession Numbers GU733320 JQ319661 FJ809784 HQ315834 EU531837 Expression analysis of each gene was confirmed in at least 3 independent RTreactions using forward and reverse primers and expression levels of each gene was normalized using actin gene as a control. Expression of genes was also analyzed by using agarose gel electrophoresis. 1.5 percent w/v agarose gel was prepared by melting 1.5 gm of agarose in 100 ml of 1 X TBE buffer in microwave oven. 4 µl (4 µl/ 100 ml) gel red solution was added to stain DNA. DNA samples were mixed with DNA loading dye containing 0.25% bromophenol blue prepared in 40% sucrose solution. Electrophoresis was performed at 100 volts (50 mA) for 65 minutes in 1 X TBE running buffer. After electrophoresis amplified product was detected by placing the gel on UV-Trans illuminator (Life Technology, USA). 4.2.3: Analysis of trichome density The number of glandular trichomes was determined at the adaxial side of 3 random pieces of fresh leaf material from each sample with an accurately determined leaf area of approximately 5mm2per leaf. Glandular trichomes were counted by using environmental scanning electron microscope (ESEM) at 200x magnification. Leaves, roots and hairy roots were cut from the transformed and non-transformed plants of Artemisia annua and Artemisia dubia using a scalpel and transferred to the microscope chamber for trichome analysis. 4.2.3.1: Environmental scanning electron microscope (ESEM) setup The sample was placed, upper epidermis upwards, on adhesive carbon tape fixed to a 10 mm stainless steel mounting disc in thermal contact with water cooled Peltier chip in the sample chamber. This allowed temperature control of the whole sample and meant that complete leaves or leaf tips could be accommodated, minimizing the number of potentially dehydrating cut edges. All experiments were conducted using an FEI XL30 Tungsten filament ESEM with the Secondary electron signal collected using a gaseous secondary electron detector. The accelerating voltage was 10-15kV. A water vapour pressure of 5mBar was used to minimize sample evaporation and samples were maintained at 2 °C. The time elapsing between cutting the leaf tissue and capturing the first image was minimized such that the tissue was as fresh as possible; this preparation and pump down time typically ranged from two to five minutes. 4.3: Results 4.3.1: Analysis of metabolic pathway and trichome development The genes ADS, CYP71AV1 and ALDH1 are involved in the metabolic synthesis pathway of artemisinin and are involved in the conversion of FDP to dihydroartemisinic acid, which is a late precursor of artemisinin (Fig. 1.15). These key genes were used to determine how transformation with the rol genes affected the pathway in the different tissues of A. annua and A. dubia. Trichome-Specific fatty acyl-CoA reductase 1(TAFR1) has been suggested to be involved in trichome development and sesquiterpenoid biosynthesis, both of which are important for artemisinin production. The expression of this gene was monitored to determine if trichome density might also be affected. 4.3.2: Relative expression of genes involved in metabolic pathway of artemisinin biosynthesis Pronounced changes in the expression of the artemisinin biosynthetic pathway genes were observed in all transgenic lines of A. annua and A. dubia transformed with rol genes (Fig. 4.1 A,B,C,D,E). The qPCR data presented here clearly demonstrate that the expression levels of ADS, CYP71AV1 and ALDH1 were significantly increased (P < 0.0001) in transformed plants of both Artemisia species compared to untransformed plants. Expression of artemisinin biosynthetic pathway genes was observed in hairy roots and roots of transformed plants compared to control roots of both A. annua and A. dubia in which low or negligible levels of expression were found. Expression levels of all the artemisinin biosynthetic genes were also significantly higher in transformed shoots compared to control plants. Interestingly transformed A. dubia showed levels, which were greater than, control A. annua plants despite similar artemisinin content. Not all the genes examined exhibited the same increase in expression. Cyp71AV1 appeared to show the greatest changes in relative expression in both shoots and hairy roots of both A. annua and A. dubia compared to the other genes. Its expression was significantly increased in transformed plants of both Artemisia species compared to untransformed plants, i.e. ~ 5 and ~ 120 times more in transformed leaves of A. annua and A. dubia, respectively. Roots of transformed plants in both Artemisia species showed significantly higher expression of this gene compared to control roots. The greatest difference in expression of this gene was found in hairy roots from both species even compared to transformed roots and control roots. Relative expression of the ADS gene was generally lower than that of the other genes however the difference in its expression compared to the control was the highest. i.e. ~ 11 and ~ 270 times greater in transformed leaves of A. annua and A. dubia respectively. Roots from these plants showed even larger differences in the expression of this gene compared to the controls. In A.annua the levels of expression of ADS was comparable to transformed roots and hairy roots but in A.dubia there was a significant difference in the expression of this gene. The ALDH1 gene showed high expression in transformed plants of both Artemisia species when compared to untransformed plants, i.e. ~ 7 and ~ 5 times more in transformed leaves of A. annua and A. dubia, respectively. Roots of these transformed plants also showed greater differences in expression of this gene in A. annua compared to control, in contrast in A.dubia the level of expression of this gene was comparatively low. However expression of this gene showed significant differences in hairy roots of both Artemisia species compared to control roots. 4.3.3: Trichome development A significant increase in expression of TFAR1 was detected in transformed plants of both Artemisia species compared to untransformed plants, i.e. ~ 10 and ~ 300 times more in transformed leaves of A. annua and A. dubia, respectively. Unexpectedly this gene was expressed in abundance in transformed roots and hairy roots of both Artemisia species. The levels of expression were greater than that seen in control shoots. Although this gene has been suggested to be responsible for trichome density it showed significant differences in expression even in transformed roots and hairy roots compared to control roots figure 4.1 (A). Relative expression of TFAR1 gene 10 9 8 7 6 5 4 3 2 1 0 Fig. 4.1 (A): Comparative analysis of expression of TFAR1 gene in shoots, roots of transformed plants and hairy roots of transformed and un-transformed plants of Artemisia dubia and Artemisia annua. The average of three plants is shown together with error bars showing SE. Relative expression of ADS gene 7 6 5 4 3 2 1 0 Fig. 4.1 (B): Comparative analysis of expression of ADS gene in shoots, roots of transformed plants and hairy roots of transformed and un-transformed plants of Artemisia dubia and Artemisia annua. The average of three plants is shown together with error bars showing SE. Relative expression of CYP71AV1 gene 12 10 8 6 4 2 0 Fig. 4.1 (C): Comparative analysis of expression of CYP71AV1 gene in shoots, roots of transformed plants and hairy roots of transformed and un-transformed plants of Artemisia dubia and Artemisia annua. The average of three plants is shown together with error bars showing SE. 10 Relative expression of ALDH1 gene 9 8 7 6 5 4 3 2 1 0 Fig. 4.1 (D): Comparative analysis of expression of ALDH1 gene in shoots, roots of transformed plants and hairy roots of transformed and un-transformed plants of Artemisia dubia and Artemisia annua. The average of three plants is shown together with error bars showing SE. 4.3.4: Statistical analysis 4.3.4.1: Comparative analysis of different genes involved in artemisinin production pathway in different tissues of Artemisia annua and Artemisia dubia Statistical analysis was conducted in factorial design (2 X 3 X 4) to see the effect of rol ABC genes on metabolic pathway artemisinin production and different genes involved in this pathway in different tissues of A.annua and A.dubia. Artemisia annua and Artemisia dubia had a significant difference (P<0.05) in production of artemisinin. Most effective results were obtained in case of different tissues of A.annua species. All genes were expressed more in A.annua plant. Table 4.4: Analysis of Variance Table for factors affecting different genes involved in artemisinin production pathway in different tissues of A.annua and A.dubia Source of Variation df Sum of Squares Species 1 61.624 Tissue Types 2 Mean F-Value Prob. 61.624 75.9609 0.0000 197.943 98.971 121.9981 0.0000 2 40.760 20.380 25.1215 0.0000 Genes 3 33.288 11.096 13.6775 0.0000 Species X Genes 3 4.236 1.412 1.7404 0.1713 6 30.181 5.030 6.2006 0.0001 6 5.571 0.928 1.1445 0.3516 Error 48 38.940 38.940 0.811 Total 71 412.542 Species X Tissue Types Tissue Types X Genes Species X Tissue Types X Genes Square Coefficient of Variation: 35.81% Effect of different tissues had significant impact (P<0.05) on production of artemisinin in A.annua and A.dubia. Mean value of artemisinin was higher in shoots of both species as compared to other tissues. Similarly, different genes of artemisinin synthesis pathway had significant impact (P<0.05) on production of artemisinin. In general CYP71AV1 gene was highly expressed as compared to other genes of pathway of artemisinin synthesis (Table 4.4). In general Artemisia annua is the best species for production of artemisinin, and among all genes expression of CYP71AV1 gene was higher in both species as compared to other genes under the effect of transformation with rol genes. Shoots are the best tissues where production of artemisinin was more enhanced compared to other tissues. Among roots, hairy roots of both species produced more artemisinin and its derivatives compared to other roots (Table 4.4). 8 B A 9 CYP ALDH1 C ADS D 6 HI EFGHI HI M M M BCD CDEF BCD L FGHI BCD CDEFG DEFG EFGHI M BCD CDEF J DEFG CDEF JK 1 EFGHI DEFG HI EFGHI 2 HI 3 CDE CDEF 4 CDE BC CDE BCDE 5 EFGHI Relative Expression of Genes 7 TFAR1 0 A.annua Transformed Shoots A.annua Control Shoots A.annua Transformed Roots A.annua Control Roots A.annua A.dubia Hairy Roots Transformed Shoots A.dubia Control Shoots A.dubia Transformed Roots A.dubia Control Roots Fig 4.2: Statistical and comparative analysis different genes involved in artemisinin production pathway in different tissues of Artemisia annua and Artemisia dubia. Each value is the mean of three replicates. Any two means having a common alphabet are not significantly different at p = 0.05 using LSD. Vertical bar represents the standard error of the 3 means. A.dubia Hairy Roots 4.3.4: Analysis of trichome density The number of glandular trichomes was determined at the adaxial side of 3 random pieces of fresh leaf material from each sample with an accurately determined leaf area of approximately 5mm2 per leaf. Glandular trichomes were counted by using environmental scanning electron microscope (ESEM) at 200x magnification. Leaves, roots and hairy roots were cut from the transformed and non-transformed plants of Artemisia annua and Artemisia dubia using a scalpel and transferred to the microscope chamber for trichome analysis. Trichome density was assessed using ESEM at low resolution and is shown in Figure 4.3 A and B. Transformed leaves of A.annua plants produced more trichomes (~ 222 trichomes/5mm2) as compared to the control leaves (~ 120 trichomes/5mm2) (Fig. 4.3A). Transformed leaves of Artemisia dubia also produced more trichomes (~ 173 trichomes/5mm2) compared to control leaves (~ 90 trichomes/5mm2). In contrast, the hairy roots, transformed and untransformed roots from both Artemisia species showed no trichome production (Fig. 4.3 A). Transformation also increased hairiness in transformed plants along with increase in trichome density compared to non transformed plants. Transformed plants produced denser network of hair on their surface compared to non-transformed plants (Fig. 4.3 B). Table.4.5: Analysis of Variance for significant difference between different species of Artemisia plant Degrees of Freedom Sum of Squares Mean Square Between 7 18652.292 2664.613 Within 16 140.667 8.792 Total 23 18792.958 Coefficient of Variation = 1.62% F-Value Prob. 303.084 0.0000 250 Number of adaxial Trichomes per 5mm2 leaf A 200 150 100 B C C D D E F 50 0 Fig. 4.3 (A): Comparative analysis of trichome density in shoots roots and hairy roots of A.annua and A.dubia. A1, A2 and A3 represents three different transgenic lines of shoots, AR1, AR2 and AR3 represents transformed roots of transgenic lines A1, A2, A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots of respectively. D1, D2 and D3 represents three transgenic lines of leaves, DR1, DR2 and DR3 represents transformed roots of transgenic lines D1, D2, D3 respectively, DH1 and DH2 represents two transgenic lines of hairy roots, DC and DRC represent control shoots and roots of respectively. The average of three plants is shown together with error bars showing SE. Data was statistically analyzed for significant difference by using ANOVA and Duncan’s Multiple Range Test. Each value is the mean of three replicates. Any two means having a different alphabet are significantly different at p = 0.05 using LSD. (A1) (A2) (A3) (A). Transformed shoots of Artemisia annua (D1) (D2) (D3) (B). Transformed Leaves of Artemisia dubia (AC) (DC) (C). Un-transformed Leaves of Artemisia annua and Artemisia dubia ( (AR1) (AR2) (AR3) (A). Transformed Roots of Artemisia annua (DR1) (DR2) (DR3) (B). Transformed Roots of Artemisia dubia (ARC) (DRC) (C). Un-transformed Roots of Artemisia annua and Artemisia dubia (AH1) (AH2) (A). Hairy Roots of Artemisia annua (DH1) (DH2) (B). Hairy Roots of Artemisia dubia Figure: 4.3 (B). Comparison of trichome density in different tissues of transformed and untransformed plants of A.annua and A.dubia. A1, A2 and A3 represents three transgenic lines of leaves in A.annua, AR1, AR2 and AR3 represents transformed roots of transgenic lines A1, A2 and A3 respectively, AH1 and AH2 represents two transgenic lines of hairy roots, AC and ARC represent control shoots and roots of A.annua respectively. D1, D2 and D3 represents three transgenic lines of leaves in A.dubia, DR1, DR2 and DR3 represents transformed roots of transgenic lines D1, D2 and D3 respectively, DH1 and DH2 represents two transgenic lines of hairy roots, DC and DRC represent control shoots and roots of A.dubia respectively. Both technical and biological replicates were performed. 4.4: Conclusion In this part, metabolic pathway of artemisinin production in Artemisia annua and Artemisia dubia and the effect of rol genes through which these genes enhance the production of artemisinin were analyzed using different genes involved in artemisinin production and TFAR1 involved in trichome development and sesquiterpenoid biosynthesis was also studied. Trichome density was also calculated in both Artemisia species and following conclusion were drawn:  Pronounced changes in the expression of the artemisinin biosynthetic pathway genes were observed in all transgenic lines of A. annua and A. dubia transformed with rol ABC genes. Expression levels of ADS, CYP71AV1 and ALDH1 were significantly increased in transformed plants and hairy roots of both Artemisia species compared to untransformed plants.  Cyp71AV1 appeared to show the greatest changes in relative expression in both shoots and hairy roots of both A. annua and A. dubia compared to the other genes. Roots of transformed plants in both Artemisia species showed significantly higher expression of this gene compared to control roots. The greatest difference in expression of this gene was found in hairy roots from both species even compared to transformed roots and control roots.  Relative expression of the ADS gene was generally lower than that of the other genes however the difference in its expression compared to the control was the highest. Roots from these plants and hairy roots showed even larger differences in the expression of this gene compared to the controls.  The ALDH1 gene showed high expression in transformed plants of both Artemisia species when compared to untransformed plants. Roots of these transformed plants also showed greater differences in expression of this gene in A. annua compared to control, in contrast in A.dubia the level of expression of this gene was comparatively low. However expression of this gene showed significant differences in hairy roots of both Artemisia species compared to control roots.  A significant increase in expression of TFAR1 was detected in transformed plants of both Artemisia species compared to untransformed plants. Unexpectedly this gene was expressed in abundance in transformed roots and hairy roots of both Artemisia species.  Transformed leaves of A.annua and A.dubia plants produced more trichomes as compared to the control leaves. In contrast, the hairy roots, transformed and untransformed roots from both Artemisia species showed no trichome production.  In general, results showed that among different genes involved in artemisinin synthesis and trichome development CYP71AV1 was highly expressed compared to other genes of metabolic pathway and among two Artemisia species, A.annua is the best species for the production of artemisinin. Considering the tissue types, shoots and hairy roots are more appropriate for production of artemisinin. Chapter No. 5 Analysis of anticancer activity on Breast Cancer Cell lines and comparison in transformed and untransformed Artemisia annua and Artemisia dubia 5.1: Introduction Cancer is a foremost cause of death all over the world. According to different data published, it is described that cancer caused 7.9 million deaths (around 13 % of all deaths) in 2007 and this number is increasing every day with an estimation of 12 million deaths in 2030. The different types of cancer which cause maximum number of deaths every year include lung, stomach, liver, colon and breast cancer. Cancer cells are produced due to changes in cell acquired by external agents or by inherited genetic factors. According to World Health Organization deaths due to cancer can be prevented up to 30% (WHO, 2008d). A large number of studies carried out by different laboratories all over the world now explain the effect of artemisinin and its derivatives on different types of cancer, such as breast cancer (Posner et al., 2004; Nam et al., 2007; Sing and Lai, 2004; Thomas et al., 2003), lung cancer (Wu et al., 2006; Thomas et al., 2003), prostate cancer (Chen et al., 2004; Thomas et al., 2003),), head and neck cancer (Yamachika et al., 2004), bladder carcinomas, renal carcinoma (Thomas et al., 2003), ovarian carcinoma (Jiao et al., 2007), cervical carcinoma (Disbrow et al., 2005), pancreas carcinoma (Wu et al., 2001), colon carcinoma (Thomas et al., 2003), thyroid medullary carcinoma, endometrial carcinoma and oral squamous cell carcinoma (Yamachika et al., 2004). To date, considerable pharmacological studies on anticancer use of artemisinin have been carried out which explain mechanism of action of artemisinin against cancer in very effective way (Efferth et al., 2001; Singh and Lai, 2001; Efferth et al., 2002; Sadava et al., 2002). According to these considerations, cancer cells require large amount of iron for their multiplication and transferrin receptors that facilitate uptake of the plasma iron-carrying protein transferrin via endocytosis, and a large concentration of cell surface to express, artemisinin reacts with iron to form free radicals which can kill cells. By covalently tagging artemisinin to transferrin, artemisinin could be selectively picked up and concentrated by cancer cells. In addition, both artemisinin and iron would be transported into the cell in one package. Once an artemisinin-tagged transferrin molecule is endocytosed, iron is released and reacts with tagged artemisinin. As a result, free radicals form that kill the cancer cells. Artemisinin-tagged transferrin is highly selective and potent in killing cancer cells. Thus, artemisinin and artemisinin-tagged iron-carrying compounds could be developed into potent anticancer drugs (Rowen, 2002; Henry et al., 2005). The aim of the investigation reported here was to analyse the anticancerous properties of transformed and non-transformed plants of Artemisia annua and Artemisia dubia. 5.2: Materials and Methods 5.2.1: Analysis of transformed and untransformed A.annua and A.dubia on breast cancer cell lines Transformed and Un-transformed plants of Artemisia annua and Artemisia dubia and hairy roots from these plants were analyzed to check their affectivity against cancer by using MCF7 breast cancer cell lines. Plants were analyzed for anticancer activity by using method of Vanicha and Kanyawim (2006) Sulforhodamine B (SRB) antiproliferative assay. Standard artemisinin was used as positive control. 5.2.2: Sample preparation Artemisinin is highly soluble in hexane/ethyl acetate while it is least soluble in water. To ensure the anticancerous activity of artemisinin, all samples were prepared in two solvents that is hexane and aqueous, because A.annua and A.dubia plants also contain secondary metabolites other than artemisinin. 5.2.2.1: Hexane extraction Biomass extraction was performed according to the method by Lapkin et al. (2006). Basically, 1g of transformed and untransformed A.annua and A.dubia plants and hairy roots were extracted with 30ml hexane/ethyl acetate (9.5:0.5) at room temperature in a Sonication bath (Kerry Ultrasonic UK) for 1 hour. Hexane was modified with ethyl acetate to increase the solubility of artemisinin in the extraction solvent. The extract was filtered over using Whatman 1.0 µm filter paper and concentrated to dryness using a rota-evaporator (Buchi R-210, USA) at 40 ◦C. Then these dried samples were dissolved in 100% (V/V) DMSO solution with 10 mg/ml for analysis on breast cancer cell lines. 5.2.2.2: Aqueous extraction 100ml of boiling water was added to 1g of transformed and untransformed A.annua and A.dubia plants and hairy roots, which was stirred and allowed to cool in the dark for 30 min, based on a modification of Van der Kooy and Verpoorte (2011) method. The filtered extract (Whatman 1 µm) was dried using a rota-evaporator (Buchi R-210, USA) at 40 ◦C. 10mg of each dried sample was dissolved in 1ml of 100% (V/V) DMSO for analysis on breast cancer lines. 5.2.3: Cell preparation 5.2.3.1: Cells The cell lines used in this study was MCF-7 human breast cancer cells. Details of its origin and characterization are described elsewhere (Sanford et al., 1948; Gey et al., 1952; Soule et al., 1973; Fogh and Trempelln, 1975). Cells grown as monolayer cultures in T-75 flasks Costar were subcultured twice a week at 37◦C in an atmosphere containing 5% CO2 in air and 100% relative humidity and maintained at low pas- sage number (5 to 20). Cells were cultured in Dulbecco‘s modified Eagle‘s medium DMEM, Warwick University, UK. Medium was supplemented with 10% fetal bovine serum (FBS), 100 µg/ml streptomycin and 100 IU/ml penicillin and additionally supplemented with 10 mg/ml insulin Sigma. 5.2.3.2: Cell inoculation Adherent cells at a logarithmic growth phase were detached by addition of 2–3 ml of a 0.05% trypsin (Gibco) 1– 0.02% EDTA mixture and incubated for 2–5 min at 37◦C. Cells were plated 200 ml per well in 96-well flat bottom microplates at densities of 1,000–100,000 cells per well. Back- ground control wells (n = 8), containing the same volume of complete culture medium, were included in each experiment. Cells were plated in sextuplicate (six replicate wells per cell density) and experiments were performed twice. Microplates were left for 6 h at 37◦C so that cells were able to attach to the bottom of the wells before the fixation protocol was carried out. 5.2.3.3: Fixation protocol In this method (Classic fixation protocol) (Skehan et al., 1990; Skehan, 1995), 50 µl of cold (4◦C) 50% TCA were added to the top of 200 µl culture medium in each well to produce a final TCA concentration of 10%. TCA was gently layered (using an 8-channel, 50 µl multipipette, tips touching the culture medium surface at the edge of the well) to avoid fluid shearing forces which could result in cell detachment and loss. Microplates were left for 30 min at 4◦C and subsequently washed 5 times with deionized water. Microplates were then dried with blow dryer. 5.2.4: Sulforhodamine B (SRB) antiproliferative assay The SRB assay was carried out as previously described (Skehan et al., 1990; Skehan, 1995). 70 µl 0.4% (w/v) Sulforhodamine B (Sigma) in 1% acetic acid solution were added to each well and left at room temperature for 20 min. SRB was removed and the plates washed 5 times with 1% acetic acid before air drying. Bound SRB was solubilized with 200 µl 10 mM unbuffered Tris-base solution (Sigma) and plates were left on a plate shaker for at least 30 min. Absorbance was read in a 96well plate reader (Anthos-2001, Anthos labteck instruments, A-5022, Salzburg) at 492 nm subtracting the background measurement at 620 nm. The test optical density (OD) value was defined as the absorbance of each individual well, minus the blank value (‗blank‘ is the mean optical density of the background control wells, n = 8). Mean values and CV from six replicate wells were calculated automatically. Curves and statistical analysis were performed using Excel 7.0 software. IC50 values were derived by using curve-fitting methods with statistical analysis software or IC50 calculation software. Percentage of control cell growth = Mean OD sample – Mean OD day 0 * 100 Mean OD negative control - mean OD day 0 Percentage growth inhibition = 100 - Percentage of control cell growth 5.3: Results 5.3.1: Analysis of transformed and untransformed Artemisia annua and Artemisia dubia on breast cancer cell lines Transformed and Un-transformed plants of Artemisia annua and Artemisia dubia and hairy roots from these plants were analyzed to check their affectivity against cancer by using MCF-7 breast cancer cell lines. Plants were analyzed for anticancer activity by using method (described in materials and methods (Section: 5.2.4)) of Vanicha and Kanyawim (2006) Sulforhodamine B (SRB) antiproliferative assay. Sulforhodamine B (SRB) antiproliferative assay was used to test anticancer activity of aqueous and hexane extracts of all transformed and untransformed plants and hairy roots of Artemisia annua and Artemisia dubia against MCF7 breast cancer cell lines at 50 µg concentration initially. The samples which showed more than 50% inhibition at 50 µg were tested further at lower concentrations i.e. at 1.0 µg and 0.1 µg to find out their IC50 values. The percentage of proliferation inhibition from transgenic lines of Artemisia annua and Artemisia dubia varied between tissues, the difference between the lines is shown (Fig. 5.1 A, B, C). The results of SRB assay showed that hexane fractions of both transformed plants of Artemisia annua and Artemisia dubia showed better proliferation inhibition than aqueous fractions against breast cancer cell lines because artemisinin is better dissolved in hexane than water which confirmed the effect of artemisinin on cancer cells compared to other secondary metabolites present in transformed and untransformed Artemisia annua and Artemisia dubia. However, the possibility of other effective compounds present in hexane extraction cannot be excluded. The hexane fractions of transformed plants of Artemisia annua showed 98% proliferation inhibition compared to 55% proliferation inhibition in the controls plants. Hairy roots of Artemisia annua also showed significantly greater (91-93%) proliferation inhibition compared to control roots in which proliferation inhibition was 51%. The hexane fraction of transformed plants of Artemisia dubia showed lower percentage of proliferation inhibition than A.annua, this was in transformed plants 93% compared to control plants 49%. Hairy roots of Artemisia dubia also showed significantly higher proliferation inhibition (85%) compared to control roots (47%) (Fig 5.1A). The aqueous fractions of transformed plants of Artemisia annua and Artemisia dubia showed lower activity as compared to hexane fractions. In aqueous fractions again transformed plants and hairy roots of Artemisia annua showed higher proliferation inhibition compared to transgenic plants of Artemisia dubia (Fig. 5.1B). The aqueous fractions of transformed plants of Artemisia annua showed 89% proliferation inhibition compared to 45% proliferation inhibition in the controls plants. Hairy roots of Artemisia annua showed significantly higher (82-84%) proliferation inhibition compared to control roots in which proliferation inhibition was 40%. The aqueous fraction of transformed plants of Artemisia dubia showed lower percentage of proliferation inhibition than A.annua, this was higher in transformed plants (84%) compared to control plants (43%). Hairy roots of Artemisia dubia also showed significantly higher proliferation inhibition (75%) compared to control roots which showed negligible proliferation inhibition (Fig 5.1B). Overall results of IC50 values showed that hexane fraction of transformed and untransformed plants of Artemisia annua and Artemisia dubia had the lowest IC50 values compared to aqueous fraction of transformed and untransformed plants of Artemisia annua and Artemisia dubia against tested MCF7 breast cancer cell lines (Fig. 5.1) because artemisinin is better dissolved in hexane than water (Lapkin and Plucinski, 2006; Wright et al., 2010 ) which confirmed the effect of artemisinin on cancer cells compared to other secondary metabolites present in transformed and untransformed Artemisia annua and Artemisia dubia. As shown in the fig 5.2, artemisinin content and anticancer activity were significantly correlated with each other in samples of hexane extracts compared to aqueous extracts. Increased anticancerous activity can also be related with the enhanced production of various secondary metabolites including artemisinin in rol genes transformed plants. Percentage Survival at 50 µg 60 50 90.01 73.6 66.6 51.41 40 30 4.71 20 2.27 10 1.3 1.75 1.84 2.43 2.01 2.53 3.03 1.96 0 Fig. 5.1 (A): Analysis of percentage survival in hexane fractions of transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia at 50 µg concentration. The IC50 values are shown as values above the bars. *Artemisinin (IC 50 = 0.21 µM) was used as positive control 140 330 Percentage Suvival at 50 µg 120 100 80 164 109 60 121.5 40 20 9.6 3 3.06 3.9 4.05 4.5 4.01 3.5 4.95 6.04 0 Fig. 5.1 (B): Analysis of percentage survival in aqueous fractions of transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia at 50 µg concentration. The IC50 values are shown as values above the bars. *Artemisinin (IC 50 = 0.21 µM) was used as positive control Hexane Fraction 120.00 Percentage Inhibition at 50 µg Aqueous Fraction 100.00 1.63 2.16 2.35 3.32 4.27 4.15 3.87 80.00 60.00 51.41 109 7.82 66.6 73.6 90.01 164 121.5 40.00 330 20.00 0.00 Fig. 5.1 (C): Comparative analysis of average percentage inhibition in hexane and aqueous fractions of transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia at 50 µg concentration. The IC50 values are shown as values above the bars. *Artemisinin (IC 50 = 0.21 µM) was used as positive control R-Squared Values 1.20 1.00 0.80 Aqeous Extraction Hexane Extraction 0.99 0.99 0.99 0.74 0.74 0.70 0.68 0.62 0.60 0.48 0.52 0.51 0.43 0.66 0.56 0.43 0.40 0.23 0.20 0.00 Fig. 5.2: Comparative correlation analysis between artemisinin content and anticancer activity in hexane and aqueous fractions of transformed and untransformed plants, roots and hairy roots of A.annua and A.dubia. The R2 values are shown as values above the bars. Those samples that had R2 value 0.60.9 were significantly correlated. Data was analyzed for significant correlation by using Graph Pad Prism Method. 5.4: Conclusion In this part, hexane and aqueous extracts from transformed and untransformed plants of Artemisia annua and Artemisia dubia were analyzed for anticancerous activity against MCF-7 breast cancer cell lines and following conclusion were drawn:  Crude extracts of rol genes transgenic plants revealed higher anticancerous activities against MCF-7 breast cancer cell compared to control plants.  Hexane extracts of rol ABC genes transgenic plants revealed higher anticancerous activity against MCF-7 breast cancer cell lines compared to aqueous extracts of plants indicating may be this activity was due to enhanced amount of artemisinin in hexane solvent.  Increased anticancerous activity could be related with the enhanced production of other secondary metabolites including artemisinin in rol transformed plants. Chapter No. 6 Sequence analysis of Artemisia annua and Artemisia dubia and comparison in transformed and untransformed plants 6.1: Introduction Plants comprise some 400,000 species and are tremendously diverse in growth habit, environmental adaptation, and nuclear genome structure. Plant genomes tend to be large and complex, varying in size from ~38 Mb (1C) for the crucifer Cardamine amara to > 87,000 Mb for Fritillaria assyriaca, a member of the Lilliaceae (Flavell et al., 1974; Bennett and Leitch, 1995). Despite this diversity, plant geneticists have found that plants exhibit extensive conservation of both gene content and gene order (Bennetzen and Freeling, 1993). Since then comparative genetic analyses have begun to show that different plant species often use homologous genes for very similar functions (Fatokun et al., 1992; Ahn et al., 1993; Paterson et al., 1995; Lagercrantz et al., 1996). Different species evolved as a result of speciation event during the course of evolution (John and Jeroen, 2004; Ohta, 1989). We come across increasingly complex genomes as we move up the evolutionary tree. There can be several reasons that have caused the complexity of gene number. Genes can be acquired from other species or they can be duplicated from the existing genes. Duplication of single gene or group of genes (Ohta, 2000), entire genome or a single chromosome or part of it can result in large genome sequences. Sequences that emerged as a result of recombination within the genome or were carried across the species by any means lead to divergence of genomes (Posada et al., 2002). Replication slippage and DNA amplification (Romero and Palacios, 1997) mostly create gene duplication. These evolutionary tools lead to sequence homology amongst different organisms. Their genomes overlap to varying degree depending upon the time of divergence. In this study, genome of A.annua and A.dubia was sequenced for the first time. Different contigs were generated by alignment of different samples individually from both A.annua and A.dubia to evaluate whether genes in the genome of transformed plants are up or down regulated compared to untransformed plants. This up and down regulation comparison was also made between A.annua and A.dubia. Furthermore, genome sequences of individual transgenic lines were also compared. BLAST of these contigs was performed using nr/nt database in NCBI. The homology of their genome sequence was found and significant results were obtained. 6.2: Materials and Methods 6.2.1: Sequence analysis of transformed and untransformed A.annua and A.dubia Sequence analysis of transformed and untransformed plants of Artemisia annua and Artemisia dubia was performed to find out differences in sequences of transformed and untransformed plants. To make libraries, 4µg of total RNA (RNA was extracted according to method described in section 4.2.1.1) was isolated and then sheared using a chemical reaction. This material was then used to create the libraries. The whole protocol is described below. 6.2.2: Purification and fragmentation of mRNA This process purifies the poly‐A containing mRNA molecules using poly‐T oligo‐attached magnetic beads using two rounds of purification. During the second elution of the poly‐A RNA, the RNA is also fragmented and primed for cDNA synthesis. 6.2.2.1: Formation and purification of RNA bead plate Following protocol was used to make and purify RBP  4µg of total RNA was diluted with nuclease‐free ultra pure water to a final volume of 50 μl in a new 96‐well 0.3 ml PCR plate with the RBP barcode label.  Thawed RNA Purification Beads tube was vigorously vortexed to completely resuspend the oligo‐dT beads.  50 μl of RNA Purification Beads were added to each well of the RBP plate using a multichannel pipette to bind the poly‐A RNA to the oligo dT magnetic beads. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column.  RBP plate was then sealed with a Microseal ‗B‘ Adhesive seal.  Sealed RBP plate was placed on the pre-programmed thermal cycler. The lid was closed and mRNA Denaturation was selected (65• ◦C for 5 minutes, 4◦C hold) to denature the RNA and facilitate binding of the poly-A RNA to the beads.  RBP plate was removed from the thermal cycler when it reached 4 ◦C.  RBP plate was placed on the bench and incubated at room temperature for 5 minutes to allow the RNA to bind to the beads.  RBP plate was placed on the magnetic stand at room temperature for 5 minutes to separate the poly -A RNA bound beads from the solution.  Adhesive seal was removed from the RBP plate.  All of the supernatant was removed and discarded from each well of the RBP plate using a multichannel pipette. Tips were changed after each column.  RBP plate was removed from the magnetic stand.  Beads were washed by adding 200 μl of Bead Washing Buffer in each well of the RBP plate using a multichannel pipette to remove unbound RNA. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column.  RBP plate was placed on the magnetic stand at room temperature for 5 minutes.  Elution Buffer was thawed and centrifuged to 600 xg for 5 seconds.  All of the supernatant was removed and discarded from each well of the RBP plate using a multichannel pipette. The tips were changed after each column.  RBP plate was removed from the magnetic stand.  50 μl of Elution Buffer was added in each well of the RBP plate using a multichannel pipette. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The RBP plate was sealed with a Microseal B Adhesive seal. The tips were changed after each column.  Elution Buffer tube was stored at 4°C.  Sealed RBP plate was placed on the pre‐programmed thermal cycler. Lid was closed and mRNA Elution 1 was selected (80°C for 2 minutes, 25°C hold) to elute the mRNA from the beads.  RBP plate was removed from the thermal cycler when it reached 25°C.  RBP plate was placed on the bench at room temperature and adhesive seal was removed from the plate. 6.2.2.2: Formation and purification of RNA fragmentation plate Following protocol was used to make and purify RFP  50 µl of Bead Binding Buffer was added to each well of the RBP plate using a multichannel pipette to allow the RNA to re-bind to the beads. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column.  RBP plate was incubated at room temperature for 5 minutes and the Bead Binding Buffer tube was stored at 2°C to 8°C.  RBP plate was placed on the magnetic stand at room temperature for 5 minutes.  All of the supernatant was removed and discarded from each well of the RBP plate using a multichannel pipette. Tips were changed after each column.  RBP plate was removed from the magnetic stand.  Beads were washed by adding 200 µl of Bead Washing Buffer in each well of the RBP plate using a multichannel pipette. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column.  Bead Washing Buffer tube was stored with at 2°C to 8°C.  RBP plate was placed on the magnetic stand at room temperature for 5 minutes.  All of the supernatant was removed and discarded from each well of the RBP plate using a multichannel pipette. Tips were changed after each column.  RBP plate was removed from the magnetic stand.  19.5 µl of Elute was added, Primed, and Fragmented and mixed to each well of the RBP plate using a multichannel pipette. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column. The Elute, Prime, Fragment Mix contains random hexamers for RT priming and serves as the 1st strand cDNA synthesis reaction buffer.  RBP plate was sealed with a Microseal ‗B‘ Adhesive seal.  Elute, Prime, Fragment Mix tube was stored at ‐15° to ‐25°C.  Sealed RBP plate was placed on the pre‐programmed thermal cycler. Lid was closed and Elution 2 ‐ Frag ‐ Prime was selected (94°C for 8 minutes, 4°C hold) to elute, fragment, and prime the RNA.  RBP plate was removed from the thermal cycler when it reaches 4°C and centrifuged briefly. 6.2.3: Synthesis of first strand cDNA This process reverse transcribes the cleaved RNA fragments primed with random hexamers into first strand cDNA using reverse transcriptase and random primers. 6.2.3.1: Formation of cDNA plate  RBP plate was placed on the magnetic stand at room temperature for 5 minutes.  Adhesive seal was removed from the RBP plate.  17 μl of the supernatant (fragmented and primed mRNA) was transferred from each well of the RBP plate to the corresponding well of the new 0.3 ml PCR plate labeled with the CDP barcode. Some liquid may remain in each well.  The thawed First Strand Master Mix was centrifuged to 600 xg for 5 seconds.  50 μl SuperScript II First Strand Master Mix tube (ratio: 1 μl SuperScript II for each 7 μl First Strand Master Mix). Mixed gently, but thoroughly, and centrifuged briefly.  First Strand Master Mix tube was labeled to indicate that the SuperScript II has been added.  8 μl of First Strand Master Mix and SuperScript II mix was added to each well of the CDP plate using a multichannel pipette. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column.  CDP plate was sealed with a Microseal ‗B‘ Adhesive seal and centrifuge briefly.  First Strand Master Mix tube back was returned to ‐15° to ‐25°C storage immediately after use.  CDP plate was incubated on the thermal cycler, with the lid closed, using the 1st Strand program:  25°C for 10 minutes  42°C for 50 minutes  70°C for 15 minutes  Hold at 4°C  When the thermal cycler reached 4°C, the CDP plate was removed from the thermal cycler. 6.2.4: Synthesis of second strand cDNA 6.2.4.1: Addition of second strand master mix  The thawed Second Strand Master Mix was centrifuged to 600 xg for 5 seconds.  Adhesive seal was removed from the CDP plate.  25 μl of thawed Second Strand Master Mix was added to each well of the CDP plate using a multichannel pipette. The entire volume was gently pipetted up and down 6 times to mix thoroughly. The tips were changed after each column.  CDP plate was sealed with a Microseal ‗B‘ Adhesive seal.  CDP plate was incubated on the pre‐heated thermal cycler; lid was closed and placed at 16°C for 1 hour.  CDP plate was removed from the thermal cycler; adhesive seal was removed, and placed the plate to room temperature.  AMPure XP beads were vortexed until they were dispersed, then 90 μl of well‐mixed AMPure XP beads were added to each well of the CDP plate containing 50 μl of ds cDNA. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  CDP plate was incubated at room temperature for 15 minutes.  CDP plate was placed on the magnetic stand at room temperature, for 5 minutes to ensure that all of the beads were bounded to the side of the wells.  135 μl of the supernatant was removed and discarded from each well of the CDP plate using a multichannel pipette. The tips were changed after each column.  200 μl of freshly prepared 80% EtOH was added to each well without disturbing the beads.  CDP plate was incubated at room temperature for 30 seconds, and then all of the supernatant was removed and discarded from each well using a multichannel pipette. The tips were changed after each column.  Steps 5 and 6 were repeated once for a total of two 80% EtOH washes.  Plate was placed at room temperature for 15 minutes to dry and then removed the CDP plate from the magnetic stand.  The thawed, room temperature Resuspension Buffer was centrifuged to 600 xg for 5 seconds.  52.5 µl OF Resuspension Buffer was added to each well of the CDP plate using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  CDP plate was incubated at room temperature for 2 minutes.  CDP plate was placed on the magnetic stand at room temperature for 5 minutes.  50 µl of the supernatant (ds cDNA) was transferred from the CDP plate to the new 0.3 ml PCR plate labeled with the IMP barcode. 6.2.5: End repairing This process converts the overhangs resulting from fragmentation into blunt ends, using an End Repair (ERP) mix. The 3' to 5' exonuclease activity of this mix removes the 3' overhangs and the polymerase activity fills in the 5' overhangs. 6.2.5.1: Formation of insert modification plate  The thawed End Repair Control tube was centrifuged to 600 xg for 5 seconds and the End Repair Control was diluted to 1/100 in Resuspension Buffer (1 μl End Repair Control + 99 μl Resuspension Buffer) before use. The diluted End Repair Control was discarded after use.  10 μl of diluted End Repair Control (or 10 μl of Resuspension Buffer if not using End Repair Control) was added to each well of the IMP plate that contains 50 μl of ds cDNA using a multichannel pipette. Tips were changed after each column.  40 μl of End Repair Mix was added to each well of the IMP plate containing the ds cDNA and End Repair Control using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  IMP plate was sealed with a Microseal ‗B‘ adhesive seal.  IMP plate was incubated on the pre‐heated thermal cycler, with the lid closed, at 30°C for 30 minutes.  IMP plate was removed from the thermal cycler.  Adhesive seal was removed from the IMP plate.  AMPure XP Beads were vortexed until they were well dispersed and 160 μl of well‐mixed AMPure XP Beads was added to each well of the IMP plate containing 100 μl of End Repair Mix. It was gently pipetted up and down 10 times to mix thoroughly.  IMP plate was incubated at room temperature for 15 minutes.  IMP plate was placed on the magnetic stand at room temperature for at least 5 minutes, until the liquid appears clear.  With 200 μl multichannel pipett set to 127.5 μl, 127.5 μl of the supernatant was removed from each well of the IMP plate. Care was taken not to disturb the beads. The tips were changed after each column.  Step 5 was repeated once.  200 μl of freshly prepared 80% EtOH was added to each well without disturbing the beads.  IMP plate was incubated at room temperature for at least 30 seconds, then removed and discarded all of the supernatant from each well. Care was taken not to disturb the beads. The tips were changed after each column.  Steps 7 and 8 were repeated once for a total of two 80% EtOH washes.  IMP plate was placed at room temperature for 15 minutes to dry and then plate was removed from the magnetic stand.  Dried pellet was resuspended in 17.5 μl Resuspension Buffer. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  IMP plate was incubated at room temperature for 2 minutes.  IMP plate was placed on the magnetic stand at room temperature for at least 5 minutes, until the liquid appeared clear.  15 μl of the clear supernatant was transferred from each well of the IMP plate to the corresponding well of the new 0.3 ml PCR plate labeled with the ALP barcode. 6.2.6: 3' Ends adenylation A single ‗A‘ nucleotide is added to the 3‘ ends of the blunt fragments to prevent them from ligating to one another during the adapter ligation reaction. A corresponding single ‗T‘ nucleotide on the 3‘ end of the adapter provides a complementary overhang for ligating the adapter to the fragment. This strategy ensures a low rate of chimera (concatenated template) formation. 6.2.6.1: Addition of A-Tailing mix  The thawed A-Tailing Control tube was centrifuged to 600 xg for 5 seconds and the A-Tailing Control was diluted to 1/100 in Resuspension Buffer (1 µl A-Tailing Control + 99 µl Resuspension Buffer) before use. The diluted ATailing Control was discarded after use.  2.5 µl of diluted A-Tailing Control (or 2.5 µl of Resuspension Buffer if not using A-Tailing Control) was added to each well of the ALP plate using a multichannel pipette.  Multichannel pipette was adjusted to 30 µl and the entire volume was gently pipetted up and down 10 times to mix thoroughly.  12.5 µl of thawed A-Tailing Mix was added to each well of the ALP plate using a multichannel pipette. Tips were changed after each column.  ALP plate was sealed with a Microseal B adhesive seal.  ALP plate was incubated on the pre-heated thermal cycler, with the lid closed, at 37°C for 30 minutes.  After 30 minutes ALP plate was removed from the thermal cycler. 6.2.7: Ligation of adapters This process ligates multiple indexing adapters to the ends of the ds cDNA, preparing them for hybridization onto a flow cell. 6.2.7.1: Addition of ligation mix  Thawed RNA Adapter Index tubes (AR001–AR012 depending on the RNA Adapter Indexes being used), Ligase Control, and Stop Ligase Mix tubes were centrifuged to 600 xg for 5 seconds.  DNA Ligase Mix tube was removed immediately from ‐15° to ‐25°C storage before use.  Adhesive seal was removed from the ALP plate.  2.5 µl of DNA Ligase Mix was added to each well of the ALP plate.  DNA Ligase Mix tube was returned back to 15 ° to 25 °C storage immediately after use.  Ligase Control was diluted to 1/100 in Resuspension Buffer (1 µl Ligase Control + 99 µl Resuspension Buffer) before use. The diluted Ligase Control was discarded after use.  2.5 μl of diluted Ligase Control (or 2.5 μl of Resuspension Buffer if not using Ligase Control) was added to each well of the ALP plate.  2.5 μl of each thawed RNA Adapter Index (AR001–AR012 depending on the RNA Adapter Indexes being used) was added to each well of the ALP plate using a multichannel pipette.  Multichannel pipette was adjusted to 40 μl and the entire volume was gently pipetted up and down 10 times to mix thoroughly.  ALP plate was sealed with a Microseal ‗B‘ adhesive seal.  ALP plate was incubated on the pre‐heated thermal cycler, with the lid closed, at 30°C for 10 minutes.  ALP plate was removed from the thermal cycler and Adhesive seal was removed from the ALP plate.  5 μl of Stop Ligase Mix was added to each well of the ALP plate to inactivate the ligation mix using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  AMPure XP Beads were vortexed until they are well dispersed, then 42 μl of mixed AMPure XP Beads was added to each well of the ALP plate using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  ALP plate was placed at room temperature for 15 minutes.  ALP plate was incubated on the magnetic stand at room temperature for at least 5 minutes, until the liquid appears clear.  79.5 μl of the supernatant from each well of the ALP plate was removed and discarded using a multichannel pipette. Some liquid may remain in each well. Tips were changed after each column.  200 μl of freshly prepared 80% EtOH was added to each well without disturbing the beads.  ALP plate was incubated at room temperature for at least 30 seconds, and then all of the supernatant was removed and discarded from each well. Care was taken not to disturb the beads. The tips were changed after each column.  Steps 5 and 6 were repeated once for a total of two 80% EtOH washes.  ALP plate was placed at room temperature for 15 minutes to dry and then plate was removed from the magnetic stand.  Dried pellet was resuspended in each well with 52.5 μl Resuspension Buffer. Gently pipetted the entire volume up and down 10 times to mix thoroughly.  ALP plate was incubated at room temperature for 2 minutes.  ALP plate was placed on the magnetic stand at room temperature for at least 5 minutes, until the liquid appeared clear.  50 μl of the clear supernatant was transferred from each well of the ALP plate to the corresponding well of the new 0.3 ml PCR plate labelled with the CAP barcode. Some liquid may remain in each well. Tips were changed after each column.  AMPure XP Beads were vortexed until they were well dispersed, then 50 μl of mixed AMPure XP Beads was added to each well of the CAP plate for a second clean up using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  CAP plate was incubated at room temperature for 15 minutes and then on the magnetic stand at room temperature for 5 minutes or until the liquid appears clear.  95 μl of the supernatant was removed and discarded from each well of the CAP plate, using a multichannel pipette. The tips were changed after each column.  200 μl of freshly prepared 80% EtOH with the CAP plate remaining on the magnetic stand was added to each well without disturbing the beads.  CAP plate was incubated at room temperature for at least 30 seconds, and then all of the supernatant was removed and discarded from each well. Care was taken not to disturb the beads. The tips were changed after each column.  Steps 17 and 18 were repeated once for a total of two 80% EtOH washes.  CAP plate was placed at room temperature for 15 minutes to dry and then removed the plate from the magnetic stand.  Dried pellet was resuspended in each well with 22.5 μl Resuspension Buffer. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  CAP plate was incubated at room temperature for 2 minutes.  CAP plate was placed on the magnetic stand at room temperature for at least 5 minutes, until the liquid appeared clear.  20 μl of the clear supernatant was transferred from each well of the CAP plate to the corresponding well of the new 0.3 ml PCR plate labeled with the PCR barcode. 6.2.8: DNA fragments enrichment This process uses PCR to selectively enrich those DNA fragments that have adapter molecules on both ends and to amplify the amount of DNA in the library. The PCR is performed with a PCR primer cocktail that anneals to the ends of the adapters. The number of PCR cycles should be minimized to avoid skewing the representation of the library. 6.2.8.1: PCR reaction  5 µl of thawed PCR Primer Cocktail was added to each well of the PCR plate using a multichannel pipette. The tips were changed after each column.  25 µl of thawed PCR Master Mix was added to each well of the PCR plate using a multichannel pipette. The tips were changed after each column. Single channel or multichannel pipette was adjusted to 40 µl and the entire volume was gently pipetted up and down 10 times to mix thoroughly. 6.2.8.2: Amplification of PCR plates  PCR plate was amplified in the pre-programmed thermal cycler, with the lid closed, using the PCR program:  98°C for 30 seconds  15 cycles of:  98°C for 10 seconds  60°C for 30 seconds  72°C for 30 seconds  72°C for 5 minutes  Hold at 4°C 6.2.8.3: Cleaning of PCR plates  AMPure XP Beads were vortexed until they were well dispersed, 50 µl of the mixed AMPure XP Beads was added to each well of the PCR plate containing 50 µl of the PCR amplified library using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  PCR plate was incubated at room temperature for 15 minutes.  PCR plate was placed on the magnetic stand at room temperature for at least 5 minutes, until the liquid appeared clear.  95 μl of the supernatant was removed and discarded from each well of the PCR plate, using a multichannel pipette. The tips were changed after each column.  200 μl of freshly prepared 80% EtOH was added to each well without disturbing the beads.  PCR plate was incubated at room temperature for at least 30 seconds, and then all of the supernatant was removed and discarded from each well. The tips were changed after each column.  Steps 5 and 6 were repeated once for a total of two 80% EtOH washes.  PCR plate was placed at room temperature for 15 minutes to dry and then removed the plate from the magnetic stand.  Dried pellet was resuspended in each well with 32.5 μl Resuspension Buffer using a multichannel pipette. The entire volume was gently pipetted up and down 10 times to mix thoroughly.  PCR plate was incubated at room temperature for 2 minutes.  PCR plate was placed on the magnetic stand at room temperature for at least 5 minutes, until the liquid appeared clear.  30 μl of the clear supernatant was transferred from each well of the PCR plate to the corresponding well of the new 0.3 ml PCR plate labeled with the TSP1 barcode. The tips were changed after each column. 6.2.9: Validation of library 6.2.9.1: Quantification of libraries In order to achieve the highest quality of data on Illumina sequencing platforms, it is important to create optimum cluster densities across every lane of every flow cell. This requires accurate Quantitation of DNA library templates. 6.2.9.2: Quality control  1 μl of the resuspended construct was loaded on an Agilent Technologies 2100 Bioanalyzer using a DNA specific chip such the Agilent DNA‐1000.  Size and purity of sample was checked. The final product should be a band at approximately 260 bp (for single‐read libraries). 6.3: Evaluation of sequence homology Different contigs were generated by alignment of different samples individually from both A.annua and A.dubia to evaluate whether genes in the genome of transformed plants are up or down regulated compared to untransformed plants. This up and down regulation comparison was also made between A.annua and A.dubia. BLAST of these contigs was performed using nr/nt database. The homology of their genome sequence was found and significant results were obtained. Putative functions of genes showing homology were sorted using Expert Protein Analysis resources. 6.4: Results In this study, the transcriptome of Artemisia annua and Artemisia dubia were sequenced for the first time. Gene contigs were generated by alignment of the sequences derived from a compilation of all the sequences obtained from a species. The different individual samples from both Artemisia annua and Artemisia dubia were then compared against this reference gene list to evaluate if any genes in the genome of the transformed plants are up or down regulated compared to untransformed plants. The Bowtie and CuffLinks programs were used to look for differential expression and CuffDiff used to derive values. A comparison was also made between Artemisia annua and Artemisia dubia (Table 6.1) plants. Contigs/genes for which significant differences were obtained were compared between biological replicates to determine their overall significance. Those which showed significant differences were submitted to the BLAST program using both the nr/nt and EST database available at NCBI. The homology of their genome sequence was found and significant results were obtained. From the results obtained the putative gene functions could be derived. Furthermore, genome sequences of individual transgenic lines, which were different from each other in copy number, were also compared. Besides, genome sequence of untransformed and transformed Artemisia annua was compared individually with genome sequence of both untransformed and transformed Artemisia dubia 6.4.1: Comparison of various transgenic lines with respect to specific contigs Transcriptome sequencing of individual A.annua and A.dubia plants allowed transcript abundance to be measured. It permitted us to determine the genes which are up or down regulated in response to transformation with the rol genes and also to determine the likelihood that they are involved directly in artemisinin production. Genes that were significantly upregulated in multiple plants compared to the controls allowed a degree of significance to be added. Comparison between A.annua and A.dubia also adds to the degree of certainty that these genes are important. As shown in Table 6.1, contigs were generated by alignment of samples in column sample_1 with samples in column sample_2 and significant value was recorded. Significant values provided information about up and down regulation of genes of respective samples. Those genes that had significant value less than 0.01 were up regulated. Genes of different transgenic lines (A1 and A2) of A.annua were highly up regulated as compared to the control plants (AC). Similar results were observed for transgenic (D1) and control (DC) lines of A. dubia (Table 6.1). When transgenic lines A1 and A2 of A.annua were compared with the transgenic line D1 of A. dubia, genes of A1 and A2 showed up regulation. Transgenic lines of A.annua were analyzed against control plants of A.dubia and genes of A.annua were found upregulated. Interesting results were seen when control lines of A. annua showed up regulation in comparison with transgenic lines of A. dubia. In this study, total number of genes identified as upregulated in each comparison produced a huge data. However, in each case, those genes that showed more than 90% homology have been shown (Table 6.1 and 6.2). Samples in column Sample_1 showed up regulation of genes as compared to the samples in column Sample_2 (Table 6.1). Number of genes that maintain 90% homology vary in each case. When A1 transgenic line was compared with A2, five genes were found upregulated. Comparison of A1 transgenic line with AC, A1 with D1, A2 with AC, and A2 with D1 showed 67, 72, 81, 63 genes upregulated respectively. 50 genes were found upregulated in comparison of the AC with D1, 28 genes of A1 with DC, 30 in A2 with DC, 55 in AC and DC and 29 genes were found upregulated in comparison of D1 with DC (Table 6.1). In some cases whole genome was found matching with the transcriptomic sequence of A.annua and A.dubia (Table 6.2). Table 6.1: Comparison of various Transgenic Lines with respect to specific Contigs S.No. Contigs Sample_1 Sample_2 Significant value 1 Contig3276:0-518 A1.1 A2.1 0.007521 2 Contig5332:0-407 A1.1 A2.1 0.00926317 3 Contig5575:0-732 A1.1 A2.1 0.0263234 4 Contig57:9-676 A1.1 A2.1 0.00230373 5 Contig9512:0-1058 A1.2 A2.1 0.0152992 6 Contig5987:0-452 A1.1 A2.2 0.00645269 7 Contig6284:0-818 A1.1 A2.2 2.93E-05 8 Contig9144:0-754 A1.1 A2.2 9.52E-07 9 Contig951:0-251 A1.1 A2.2 0.0199822 10 Contig100:43-844 A1.1 AC.1 8.10E-05 11 Contig10524:0-757 A1.1 AC.1 0.000706668 12 Contig10536:0-983 A1.1 AC.1 0.00529889 13 Contig1181:0-4663 A1.1 AC.1 0.00408348 14 Contig1260:39-212 A1.1 AC.1 0.00822729 15 Contig1320:0-1308 A1.1 AC.1 0.0272819 16 Contig1606:21-1047 A1.1 AC.1 0.00080821 17 Contig161:0-979 A1.1 AC.1 4.06E-06 18 Contig167:6-215 A1.1 AC.1 0.0397222 19 Contig1746:0-2322 A1.1 AC.1 0.0184849 20 Contig2006:0-880 A1.1 AC.1 0.00328411 21 Contig2028:0-994 A1.1 AC.1 0.0489583 22 Contig2219:42-1241 A1.1 AC.1 0.00549141 23 Contig2799:0-245 A1.1 AC.1 0.00353468 24 Contig2903:0-2391 A1.1 AC.1 0.0170782 25 Contig8369:11-1513 A1.2 AC.1 4.68E-07 26 Contig2157:15-764 A1.1 AC.2 2.29E-07 27 Contig5597:0-1737 A1.1 AC.2 8.97E-08 28 Contig2380:0-1787 A1.2 AC.2 5.24E-12 29 Contig231:0-650 A1.2 AC.1 0.000630584 30 Contig249:21-956 A1.2 AC.1 0.000107155 31 Contig253:0-1100 A1.2 AC.1 0.000110396 32 Contig2799:0-245 A1.2 AC.1 0.000816532 33 Contig2859:34-864 A1.2 AC.1 0.000625534 34 Contig3121:0-262 A1.2 AC.1 0.00026362 35 Contig2157:15-764 A2.1 AC.2 2.90E-09 36 Contig8369:11-1513 A2.1 AC.1 5.05E-08 37 Contig2380:0-1787 A2.2 AC.2 1.90E-08 38 Contig5597:0-1737 A2.2 AC.2 4.96E-08 39 Contig10105:0-1419 A2.1 AC.2 6.57E-05 40 Contig161:0-979 A2.1 AC.2 9.64E-10 41 Contig3836:0-1711 A2.1 AC.2 5.10E-08 42 Contig3946:2-446 A2.1 AC.2 3.66E-08 43 Contig515:13-967 A2.1 AC.2 3.61E-11 44 Contig8590:0-1715 A2.1 AC.2 1.35E-08 45 Contig8696:0-1843 A2.1 AC.2 1.42E-08 46 Contig3836:0-1711 A2.2 AC.2 3.61E-11 47 Contig3840:0-1556 A2.2 AC.2 5.51E-08 48 Contig8590:0-1715 A2.2 AC.2 1.83E-08 49 Contig1972:29-1001 A2.1 AC.2 4.10E-05 50 Contig231:0-650 A2.1 AC.2 9.41E-06 51 Contig2622:14-797 A2.1 AC.2 3.38E-05 52 Contig3934:0-1618 A2.1 AC.2 9.68E-05 53 Contig4225:0-563 A2.1 AC.2 3.50E-05 54 Contig4679:38-421 A2.1 AC.2 1.59E-05 55 Contig5845:0-796 A2.1 AC.2 1.24E-05 56 Contig981:16-763 A2.1 AC.2 2.25E-05 57 Contig10105:0-1419 A2.2 AC.2 8.41E-05 58 Contig1907:54-728 A2.2 AC.2 3.83E-05 59 Contig231:0-650 A2.2 AC.2 9.02E-06 60 Contig3934:0-1618 A2.2 AC.2 3.74E-05 61 Contig4225:0-563 A2.2 AC.2 8.01E-05 62 Contig436:0-616 A2.2 AC.2 1.50E-06 63 Contig4679:38-421 A2.2 AC.2 1.50E-05 64 Contig4749:0-1057 A2.2 AC.2 3.82E-05 65 Contig8881:0-761 A2.2 AC.2 3.52E-06 66 Contig981:16-763 A2.2 AC.2 1.38E-05 67 Contig10105:0-1419 A1.1 D1.1 1.78E-06 68 Contig1388:5-1056 A1.1 D1.1 2.10E-05 69 Contig296:0-295 A1.1 D1.1 1.19E-06 70 Contig3057:0-1927 A1.1 D1.1 8.73E-12 71 Contig3437:0-1709 A1.1 D1.1 1.75E-10 72 Contig4685:0-615 A1.1 D1.1 5.65E-05 73 Contig5168:6-688 A1.1 D1.1 0.000201614 74 Contig5601:0-759 A1.1 D1.1 2.04E-07 75 Contig7008:0-1433 A1.1 D1.1 1.21E-06 76 Contig7150:0-1455 A1.1 D1.1 1.50E-05 77 Contig718:0-1984 A1.1 D1.1 2.01E-10 78 Contig7416:0-1030 A1.1 D1.1 0.000153826 79 Contig7540:0-1080 A1.1 D1.1 2.91E-05 80 Contig8220:0-844 A1.1 D1.1 1.59E-05 81 Contig10057:0-852 A1.2 D1.1 0.00151277 82 Contig10105:0-1419 A1.2 D1.1 2.23E-06 83 Contig1388:5-1056 A1.2 D1.1 1.70E-05 84 Contig1972:29-1001 A1.2 D1.1 4.96E-08 85 Contig209:0-247 A1.2 D1.1 0.000277453 86 Contig2454:22-1644 A1.2 D1.1 0 87 Contig249:21-956 A1.2 D1.1 0.000363911 88 Contig2799:0-245 A1.2 D1.1 0.000286909 89 Contig296:0-295 A1.2 D1.1 1.26E-06 90 Contig3057:0-1927 A1.2 D1.1 8.73E-12 91 Contig3437:0-1709 A1.2 D1.1 2.47E-10 92 Contig3946:2-446 A1.2 D1.1 8.57E-09 93 Contig5601:0-759 A1.2 D1.1 2.16E-07 94 Contig7008:0-1433 A1.2 D1.1 1.42E-06 95 Contig7150:0-1455 A1.2 D1.1 1.50E-05 96 Contig718:0-1984 A1.2 D1.1 1.50E-10 97 Contig7469:34-524 A1.2 D1.1 5.72E-05 98 Contig5597:0-1737 A1.2 D1.2 8.73E-12 99 Contig10057:0-852 A2.1 D1.1 3.16E-05 100 Contig10105:0-1419 A2.1 D1.1 2.46E-07 101 Contig1388:5-1056 A2.1 D1.1 1.88E-07 102 Contig1972:29-1001 A2.1 D1.1 4.96E-08 103 Contig2698:0-454 A2.1 D1.1 1.10E-07 104 Contig2867:23-797 A2.1 D1.1 1.10E-07 105 Contig296:0-295 A2.1 D1.1 5.46E-06 106 Contig3946:2-446 A2.1 D1.1 8.72E-09 107 Contig4155:0-1303 A2.1 D1.1 1.79E-05 108 Contig451:21-2271 A2.1 D1.1 2.16E-05 109 Contig4685:0-615 A2.1 D1.1 2.73E-05 110 Contig478:36-283 A2.1 D1.1 5.33E-05 111 Contig5575:0-732 A2.1 D1.1 2.52E-05 112 Contig5601:0-759 A2.1 D1.1 3.87E-09 113 Contig7008:0-1433 A2.1 D1.1 1.83E-08 114 Contig7015:0-567 A2.1 D1.1 7.88E-07 115 Contig7150:0-1455 A2.1 D1.1 1.33E-05 116 Contig718:0-1984 A2.1 D1.1 4.32E-11 117 Contig7219:0-736 A2.1 D1.1 2.93E-05 118 Contig10057:0-852 A2.2 D1.1 3.33E-05 119 Contig10105:0-1419 A2.2 D1.1 3.83E-07 120 Contig1388:5-1056 A2.2 D1.1 2.82E-07 121 Contig2698:0-454 A2.2 D1.1 1.07E-07 122 Contig2867:23-797 A2.2 D1.1 1.10E-07 123 Contig296:0-295 A2.2 D1.1 3.22E-06 124 Contig5601:0-759 A2.2 D1.1 2.90E-09 125 Contig5987:0-452 A2.2 D1.1 1.14E-08 126 Contig6284:0-818 A2.2 D1.1 1.71E-06 127 Contig718:0-1984 A2.2 D1.1 4.96E-11 128 Contig7219:0-736 A2.2 D1.1 3.19E-05 129 Contig5597:0-1737 A2.2 D1.2 5.24E-12 130 Contig5597:0-1737 A2.1 D1.2 0 131 Contig10461:0-582 AC.1 D1.1 4.88E-05 132 Contig2006:0-880 AC.1 D1.1 5.29E-07 133 Contig2126:0-1124 AC.1 D1.1 4.66E-11 134 Contig2288:0-726 AC.1 D1.1 1.20E-09 135 Contig231:0-650 AC.1 D1.1 4.40E-10 136 Contig2454:22-1644 AC.1 D1.1 4.48E-09 137 Contig2622:14-797 AC.1 D1.1 7.33E-06 138 Contig290:0-475 AC.1 D1.1 3.12E-09 139 Contig318:0-498 AC.1 D1.1 1.94E-05 140 Contig362:0-275 AC.1 D1.1 7.85E-06 141 Contig4168:0-828 AC.1 D1.1 1.71E-06 142 Contig418:2-2747 AC.1 D1.1 5.65E-05 143 Contig4225:0-563 AC.1 D1.1 1.66E-05 144 Contig4704:0-1313 AC.1 D1.1 2.65E-09 145 Contig5025:0-1950 AC.1 D1.1 6.92E-10 146 Contig5482:0-902 AC.1 D1.1 2.78E-07 147 Contig583:0-1445 AC.1 D1.1 3.77E-05 148 Contig6601:0-889 AC.1 D1.1 7.93E-07 149 Contig762:0-855 AC.1 D1.1 8.72E-07 150 Contig8568:0-1291 AC.1 D1.1 9.39E-05 151 Contig2380:0-1787 AC.2 D1.1 0 152 Contig4704:0-1313 AC.2 D1.1 3.26E-09 153 Contig4971:0-638 AC.2 D1.1 9.95E-07 154 Contig5008:0-2908 AC.2 D1.1 2.66E-08 155 Contig10081:0-821 A1.1 DC.1 3.40E-08 156 Contig10271:0-621 A1.1 DC.1 1.22E-05 157 Contig1852:0-2203 A1.1 DC.1 0 158 Contig2157:15-764 A1.1 DC.1 3.50E-08 159 Contig2380:0-1787 A1.1 DC.1 0.0092017 160 Contig3057:0-1927 A1.1 DC.1 2.19E-08 161 Contig5601:0-759 A1.1 DC.1 1.90E-05 162 Contig5821:0-949 A1.1 DC.1 5.32E-06 163 Contig7708:0-1368 A1.1 DC.1 4.88E-05 164 Contig9460:0-1091 A1.1 DC.1 8.06E-05 165 Contig9658:53-1094 A1.1 DC.1 5.85E-05 166 Contig10081:0-821 A1.2 DC.1 1.90E-08 167 Contig1852:0-2203 A1.2 DC.1 0 168 Contig1972:29-1001 A1.2 DC.1 7.11E-06 169 Contig2157:15-764 A1.2 DC.1 3.79E-08 170 Contig2380:0-1787 A1.2 DC.1 0.00925958 171 Contig3057:0-1927 A1.2 DC.1 1.84E-08 172 Contig5601:0-759 A1.2 DC.1 1.97E-05 173 Contig5821:0-949 A1.2 DC.1 3.71E-06 174 Contig9460:0-1091 A1.2 DC.1 6.54E-05 175 Contig9658:53-1094 A1.2 DC.1 7.28E-05 176 Contig10867:0-1043 A1.1 DC.1 0.000300206 177 Contig3512:0-616 A1.1 DC.1 0.0020524 178 Contig5184:30-251 A1.1 DC.1 0.000127465 179 Contig10271:0-621 A1.2 DC.1 0.000345329 180 Contig1514:18-187 A1.2 DC.1 0.00957196 181 Contig851:0-1951 A1.2 DC.1 0.000573248 182 Contig10081:0-821 A2.1 DC.1 5.97E-08 183 Contig161:0-979 A2.1 DC.1 5.85E-05 184 Contig1852:0-2203 A2.1 DC.1 0 185 Contig1972:29-1001 A2.1 DC.1 7.14E-06 186 Contig2157:15-764 A2.1 DC.1 4.11E-10 187 Contig3057:0-1927 A2.1 DC.1 1.37E-09 188 Contig5601:0-759 A2.1 DC.1 1.37E-06 189 Contig5624:0-2559 A2.1 DC.1 0.00684456 190 Contig5821:0-949 A2.1 DC.1 2.92E-05 191 Contig584:0-908 A2.1 DC.1 0.00560855 192 Contig7019:0-704 A2.1 DC.1 4.54E-05 193 Contig7251:0-1059 A2.1 DC.1 3.74E-05 194 Contig7708:0-1368 A2.1 DC.1 1.98E-05 195 Contig851:0-1951 A2.1 DC.1 0.000457521 196 Contig9460:0-1091 A2.1 DC.1 0.000428194 197 Contig10081:0-821 A2.2 DC.1 6.10E-08 198 Contig161:0-979 A2.2 DC.1 6.25E-05 199 Contig1852:0-2203 A2.2 DC.1 0 200 Contig1907:54-728 A2.2 DC.1 3.91E-06 201 Contig2139:0-1197 A2.2 DC.1 0.0329254 202 Contig2157:15-764 A2.2 DC.1 4.70E-10 203 Contig3057:0-1927 A2.2 DC.1 1.20E-09 204 Contig359:18-539 A2.2 DC.1 0.000271045 205 Contig4155:0-1303 A2.2 DC.1 0.00146942 206 Contig451:21-2271 A2.2 DC.1 2.48E-05 207 Contig462:0-927 A2.2 DC.1 6.10E-08 208 Contig7251:0-1059 A2.2 DC.1 1.64E-05 209 Contig7708:0-1368 A2.2 DC.1 2.81E-05 210 Contig851:0-1951 A2.2 DC.1 0.000567366 211 Contig2006:0-880 AC.1 DC.1 2.14E-05 212 Contig231:0-650 AC.1 DC.1 4.06E-06 213 Contig3946:2-446 AC.1 DC.1 4.75E-06 214 Contig418:2-2747 AC.1 DC.1 0.00570961 215 Contig4225:0-563 AC.1 DC.1 0.00025784 216 Contig4704:0-1313 AC.1 DC.1 1.03E-06 217 Contig489:21-1070 AC.1 DC.1 0.00109131 218 Contig7251:0-1059 AC.1 DC.1 3.85E-07 219 Contig7255:2-1153 AC.1 DC.1 0.00249426 220 Contig7356:0-819 AC.1 DC.1 1.54E-06 221 Contig8369:11-1513 AC.1 DC.1 0.000153811 222 Contig8590:0-1715 AC.1 DC.1 2.26E-05 223 Contig8879:0-594 AC.1 DC.1 5.61E-05 224 Contig8881:0-761 AC.1 DC.1 9.90E-07 225 Contig8890:0-844 AC.1 DC.1 1.46E-06 226 Contig1845:0-590 AC.2 DC.1 7.03E-05 227 Contig2006:0-880 AC.2 DC.1 5.08E-05 228 Contig231:0-650 AC.2 DC.1 9.69E-06 229 Contig2380:0-1787 AC.2 DC.1 0.00027316 230 Contig3946:2-446 AC.2 DC.1 3.72E-06 231 Contig4225:0-563 AC.2 DC.1 0.000764351 232 Contig4704:0-1313 AC.2 DC.1 1.22E-06 233 Contig7019:0-704 AC.2 DC.1 0.000754809 234 Contig7251:0-1059 AC.2 DC.1 9.90E-07 235 Contig7356:0-819 AC.2 DC.1 4.73E-06 236 Contig8590:0-1715 AC.2 DC.1 6.97E-05 237 Contig8879:0-594 AC.2 DC.1 0.000120113 238 Contig8881:0-761 AC.2 DC.1 1.75E-06 239 Contig8890:0-844 AC.2 DC.1 3.19E-06 240 Contig8369:11-1513 AC.1 DC.2 0.000128175 241 Contig2380:0-1787 D1.1 DC.2 0.00154775 242 Contig2380:0-1787 D1.1 DC.1 0.00253405 243 Contig2867:23-797 D1.1 DC.1 5.25E-08 244 Contig3395:0-424 D1.1 DC.1 5.29E-08 245 Contig3512:0-616 D1.1 DC.1 7.81E-07 246 Contig362:0-275 D1.1 DC.1 3.74E-05 247 Contig3946:2-446 D1.1 DC.1 1.42E-06 248 Contig6427:0-816 D1.1 DC.1 2.00E-05 249 Contig6578:0-492 D1.1 DC.1 5.05E-05 250 Contig7008:0-1433 D1.1 DC.1 0.000150039 251 Contig7015:0-567 D1.1 DC.1 0.000161943 252 Contig7150:0-1455 D1.1 DC.1 2.97E-06 253 Contig7546:0-614 D1.1 DC.1 2.39E-05 254 Contig7708:0-1368 D1.1 DC.1 0.000175604 255 Contig848:0-283 D1.1 DC.1 2.87E-05 256 Contig8568:0-1291 D1.1 DC.1 1.58E-06 257 Contig2380:0-1787 D1.2 DC.1 0.00214506 258 Contig2867:23-797 D1.2 DC.1 7.11E-08 259 Contig3395:0-424 D1.2 DC.1 5.28E-08 260 Contig3512:0-616 D1.2 DC.1 6.17E-07 261 Contig362:0-275 D1.2 DC.1 3.24E-05 262 Contig3946:2-446 D1.2 DC.1 1.38E-06 6.4.2: Contigs of A.annua and A.dubia producing more than 90% homology This study of transcriptome of A. annua and A. dubia first time allowed sequence of functional genes. On the basis of established fact of evolution from common ancestor and sequence conservation, this new transcriptomic sequence was matched with genes with known functions and sequences in nr/nt database to predict their putative molecular functions. Additionally, structures of protein and different kind of ribonucleic acids can also be predicted for these newly sequenced genes. Evolutionarily close relatives of A. annua and A. dubia can be proposed through phylogenetic tree construction. Percentage similarity gives the idea about how far the two species have lived in genetic time period. In order to validate the sequences, BLAST of contigs, obtained through alignment of sequences of different transcriptomic sequences of these Artemisia species, was performed. 16400 contigs were generated by alignment of different samples of A. annua and A. dubia. Only 500 contigs were selected for BLAST on the basis of significant value and in order to properly manage data for the study. Of these 500 contigs, 264 showed homology with genes of different organisms. Only those contigs were selected for analysis and putative gene functions that showed more than 90% homology. BLAST results of Contig3276:0-518 showed that gene WRKY of Malus x domestica has percentage similarity of 94%. This gene is a transcription factor that plays a key role in artemisinin synthesis pathway in Artemisia. Similarly, BLAST results of contigs Contig5597:0-1737 showed that genes CBS, CBS732, CBS6284 of Naumovozyma dairenensis, Zygosaccharomyces rouxii, and Tetrapisispora blattae respectively, Contig10105:0-1419 with gene CCPU of Cynara cardunculus and Contig3946:2-446 with gene TX077308 of Francisella sp. have percentage similarity 92-100%. These genes are related to cytochrome that plays an important role in artemisinin biosynthesis pathway. Sequences of contigs which showed 90% or more homology were used for BLAST and for other studies are given in appendix VI. Table 6.2: Sequences producing significant Homology S.No . Contigs Source Gene Putative Gene Function Accession No. Transcription factor HM122720.1 E-Value %age Similarity A1 AND A2 1 2 3 Contig3276:0518 Malus x domestica Contig57:9-676 Lactuca sativa Contig5987:0452 Solanum lycopersicum Solanum lycopersicum Lycopersicon esculentum WRKY17 psaH Docking of the LHC I antenna complex LEFL1016AF02 rRNA binding FC11AF10 134016F NADH dehydrogenase activity RNA-directed DNA polymerase activity 0.001 94% AF162205.1 4.00E102 91% AK321003.1 2.00E-13 93% AK246443.1 2.00E-13 93% BT014562.1 1.00E-10 92% A1 AND AC 5 6 Contig100:43844 Solanum nigrum Metallothioneinlike mRNA Metal ion binding FJ402840.1 4.00E-07 95% Flaveria bidentis pNDK-b ATP-binding, Kinase Activity U10283.1 1.00E167 90% NDK ATP-binding, Kinase Activity U72142.1 8.00E164 90% Contig1606:211047 Helianthus annuus 7 8 10 Contig161:0979 Contig167:6215 Contig2157:15764 Brassica oleracea DQ267199.1 6.00E-64 90% 95% Flaveria anomala gdcsPA Oxidoreductase Z36879.1 Flaveria anomala gdcsP Oxidoreductase Z99762.1 0.00E+0 0 95% Flaveria anomala gdcsPB gene Oxidoreductase Z54239.1 1.00E-64 95% Vitis vinifera ENTAV Signal transducer activity AM485693.1 8.00E-41 91% Ajellomyces capsulatus HCAG_07571 Decarboxylating and phosphate binding XM_001537 212.1 1.00E-08 94% Artemisia annua CAZI5271 Structural constituent of ribosome EY083474.1 0.00E+0 0 95% Artemisia annua CAZI5271 Structural constituent of ribosome EY107471.1 7.00E-31 100% Coffea arabica CA00-XX-FB2022-A07-JE Nucleotidyltransferase GT702383.1 1.00E-04 97% Koi herpesvirus Contig5597:01737 CA-responsive protein 0.00E+0 0 Cyprinid herpesvirus 11 BoCAR44 Naumovozyma dairenensis Zygosaccharomyce s rouxii FL ORF148 Membrane protein ORF148 GU815504.1 3.00E-07 97% KHV-I Serine-type endopeptidase activity DQ177346.1 3.00E-07 100% CBS Activator of cytochrome b HE580273.1 1.00E-06 97% CBS732 Activator of cytochrome b CU928173.1 1.00E-06 97% Tetrapisispora blattae Kuenenia stuttgartiensis CBS 6284 Activator of cytochrome b HE806321.1 1.00E-05 97% KUST_C Repressor CT573073.1 1.00E-05 97% Pantoea ananatis PA13 Catalytic activity CP003085.1 2.00E-04 100% Thielavia terrestris NRRL 8126 Shikimate kinase activity CP003011.1 2.00E-04 97% Zygosaccharomyce s rouxii CBS732 MAP kinase activity CU928175.1 2.00E-04 100% Neurospora crassa B11B23 Pyridoxal phosphate binding AL669991.1 2.00E-04 100% Trypanosoma brucei gambiense DAL972 Endonuclease activity FN554971.1 6.00E-04 100% Capsicum annuum CaCM529L23 Vasodilation GU048888.1 0.002 100% C08HBa0210J04 Exonuclease AP009282.1 0.002 100% CgPTK-l ATP binding AB098199.1 4.00E-06 97% DDB_G0280785 DNA binding 4.00E-06 97% Puccinia graminis ketoacyl-CoA thiolase A 2.00E-04 100% Solanum lycopersicum LEFL1039CB09 Acetyl-CoA CAcyltransferase activity Receptor tyrosine-protein kinase XM_635935. 1 XM_003329 597.1 AK322583.1 6.00E-04 94% Drosophila virilis Dvir\GJ13854 Complete genome 6.00E-04 94% Arabidopsis thaliana IKU1 Regulation of seed growth XM_002048 196.1 NM_129075. 3 0.002 100% Solanum lycopersicum Codonosiga gracilis Dictyostelium discoideum 12 Contig2380:01787 EY104210.1 6.00E133 96% Artemisia annua CAZI3603 Structural constituent of ribosome Lotus japonicus BP056955 DNA binding GO022661.1 2.00E-08 97% Solanum tuberosum POADE79 Trichodiene synthase activity CK274023.1 5.00E-05 97% Pinus pinaster PPSW21 Oxidoreductase BX000758.1 5.00E-05 95% Honey bee BH10057G22 RNA binding DB744146.1 2.00E-04 100% Cynara cardunculus CCPX3860 Serine-type endopeptidase activity GE606757.1 6.00E-04 94% Artemisia annua CAZI23588 Structural constituent of ribosome EY080578.1 0.00E+0 0 98% 91% Lactuca virosa CLVY6837 Calcium ion binding DW168485.1 0.00E+0 0 Centaurea solstitialis CNSM11838 ATP Binding EH765839.1 0.00E+0 0 92% 91% Taraxacum officinale CTOY4902 DNA binding DY829234.1 0.00E+0 0 Centaurea maculosa CNMM5610 DNA binding EH732731.1 0.00E+0 0 91% Helianthus exilis CHEM3111 Transducer EE643217.1 0.00E+0 90% 0 13 14 Contig249:21956 Contig253:01100 Contig3121:0262 15 3.00E-15 92% Structural constituent of ribosome XM_003608 228.1 XM_001265 256.1 1.00E-13 93% ATCC 42464 Hydrolase activity CP003005.1 2.00E-16 95% Pisum sativum PUB2 Ubiquitin-protein ligase activity L81140.1 4.00E-14 92% Ricinus communis FBA, putative mRNA Hydro-lyase activity XM_002528 547.1 1.00E-06 90% Rattus norvegicus Col18a1 Structural molecule activity 2.00E-06 Drosophila virilis Dvir\GJ16689 Complete genome NM_053489. 2 XM_002056 755.1 2.00E-06 97% Zea mays clone EL01N0361G07. c rRNA binding BT017098.1 2.00E-06 97% Drosophila persimilis Dper\GL14534 Zinc ion binding 8.00E-06 97% Drosophila virilis Dvir\GJ18864 Complete genome 1.00E-04 100% Drosophila erecta Dere\GG17580 Complete genome 1.00E-04 97% Oryza sativa Os06g0589800 ATP Binding XM_002022 800.1 XM_002055 268.1 XM_001978 582.1 NM_001064 499.1 1.00E-04 95% Mus musculus RP23-365N23 RNA binding AC159893.2 5.00E-08 95% Medicago truncatula Neosartorya fischeri Myceliophthora thermophila MTR_4g091580 PTM NFIA_020680 95% Mouse DNA RP23-123H11 RNA binding CT030637.9 2.00E-07 95% Rattus norvegicus RNECO-134C12 DNA binding AC245859.6 2.00E-06 97% Candida parapsilosis Human chromosome Drosophila melanogaster CDC317 Oxidoreductase HE605203.1 2.00E-06 97% BAC R-909M7 Antibiotic AL132709.5 2.00E-06 95% BACR-32D05 Antibiotic 8.00E-06 100% Mouse DNA RP23-80C18 RNA binding AC010212.1 0 AL590864.1 2 8.00E-06 97% ATCC 42464 Hydrolase activity CP003007.1 3.00E-05 100% 384-01M13F.t2F_J01 LjT02G13, TM0489a Sequence-specific DNA binding HQ437322.1 1.00E-04 95% DNA binding AP009827.1 1.00E-04 97% Otocolobus manul SRY Activator, Repressor DQ095178.1 1.00E-04 97% Echinometra sp. T-Ftm5_A1 Binding Function AY261271.1 1.00E-04 100% Artemisia annua CAZI5271 Structural constituent of ribosome EY083474.1 0.00E+0 0 95% Artemisia annua CAZI5271 Structural constituent of ribosome EY107471.1 7.00E-31 100% Myceliophthora thermophila Alexandrium ostenfeldii Lotus japonicus A2 AND AC 16 Contig2157:15764 17 18 19 Contig8369:111513 Contig2380:01787 Contig5597:01737 Coffea arabica CA00-XX-FB2022-A07-JE Artemisia annua CTOX5296.b1_ P04 Artemisia annua Lactuca virosa GT702383.1 1.00E-04 97% Integral to membrane EY115636.1 1.00E163 95% CAZI23588 Structural constituent of ribosome EY080578.1 0.00E+0 0 98% CLVY6837 Calcium ion binding DW168485.1 0.00E+0 0 91% 92% RNA binding Centaurea solstitialis CNSM11838 ATP Binding EH765839.1 0.00E+0 0 Taraxacum officinale CTOY4902 DNA binding DY829234.1 0.00E+0 0 91% 91% 90% Centaurea maculosa CNMM5610 DNA binding EH732731.1 0.00E+0 0 Helianthus exilis CHEM3111 Transducer EE643217.1 0.00E+0 0 Cyprinid herpesvirus FL ORF148 Membrane protein ORF148 GU815504.1 Koi herpesvirus KHV-I ATP Binding DQ177346.1 3.00E-07 100% Naumovozyma dairenensis CBS Translation regulator activity HE580273.1 1.00E-06 97% 3.00E-07 97% Zygosaccharomyce s rouxii Tetrapisispora blattae Kuenenia stuttgartiensis CBS732 Translation regulator activity CU928173.1 1.00E-06 97% CBS 6284 Translation regulator activity HE806321.1 1.00E-05 97% KUST_C Repressor CT573073.1 1.00E-05 97% Pantoea ananatis PA13 Catalytic activity CP003085.1 2.00E-04 100% Thielavia terrestris NRRL 8126 Shikimate kinase activity CP003011.1 2.00E-04 100% Zygosaccharomyce s rouxii CBS732 Translation regulator activity CU928175.1 2.00E-04 100% Neurospora crassa B11B23 Pyridoxal phosphate binding AL669991.1 2.00E-04 100% Trypanosoma brucei gambiense DAL972 Endonuclease activity FN554971.1 6.00E-04 100% Capsicum annuum CaCM529L23 Vasodilation GU048888.1 0.002 100% C08HBa0210J04 Exonuclease AP009282.1 0.002 100% CgPTK-l ATP binding AB098199.1 4.00E-06 97% DDB_G0280785 DNA binding 4.00E-06 97% Puccinia graminis ketoacyl-CoA thiolase A 2.00E-04 100% Solanum lycopersicum LEFL1039CB09 Acetyl-CoA CAcyltransferase activity Receptor tyrosine-protein kinase XM_635935. 1 XM_003329 597.1 AK322583.1 6.00E-04 94% Drosophila virilis Dvir\GJ13854 XM_002048 196.1 6.00E-04 94% Solanum lycopersicum Codonosiga gracilis Dictyostelium discoideum Complete genome NM_129075. 3 0.002 100% CAZI3603 Structural constituent of ribosome EY104210.1 6.00E133 96% Lotus japonicus BP056955 DNA binding GO022661.1 Solanum tuberosum POADE79 Trichodiene synthase activity CK274023.1 5.00E-05 97% Pinus pinaster PPSW21 Oxidoreductase BX000758.1 5.00E-05 95% Honey bee BH10057G22 RNA binding DB744146.1 2.00E-04 100% Cynara cardunculus CCPX3860 GE606757.1 6.00E-04 94% Artemisia annua CAZI3763.CAZI EY104502.1 0 99% Cynara cardunculus CCPU963 Heme binding, cytochrome c GE585313.1 2.00E-15 92% Brassica oleracea BoCAR44 CA-responsive protein DQ267199.1 6.00E-64 90% Drosophila yakuba Dyak\GE24309 Complete proteome XM_002097 700.1 4.00E-04 95% Francisella sp. TX077308 Complete proteome CP002872.1 0.001 100% KLTH0F07854g Zinc ion binding XM_002554 498.1 0.001 100% CBS 6340 Translation regulator activity CU928170.1 0.001 100% Arabidopsis thaliana 20 21 22 Contig10105:01419 Contig161:0979 Contig3946:2446 IKU1 Regulation of seed growth Artemisia annua Lachancea thermotolerans Lachancea thermotolerans Serine-type endopeptidase activity Structural constituent of ribosome 2.00E-08 97% XM_002138 246.1 XM_002054 269.1 XM_002023 391.1 0.001 100% 0.001 100% 0.001 100% Chaperone CP000937.1 0.001 92% IP09721 Sequence-specific DNA binding BT029275.1 0.001 100% Eriocheir sinensis Esc29 Repressor DQ450915.1 0.001 100% Roridula dentata CONN12998 DNA binding JQ519390.1 3.00E-36 97% Asclepias subulata OKLA JN665094.1 3.00E-36 97% Daucus carota BAC C237E06 FJ150367.1 3.00E-36 97% Camellia sinensis XUTDF2 DNA binding EU730587.1 3.00E-36 97% Solanum lycopersicum hba-17c15 map 1 Serine/threonine-protein kinase AC246154.3 5.00E-96 90% Candida glabrata CBS138 Translation regulator activity CR380959.2 4.00E-09 92% Clavispora lusitaniae ATCC 42720 Chaperone XM_002619 926.1 8.00E-11 94% Homo sapiens PRKCE ATP binding EU332867.1 0.005 100% Zebrafish CH73-269F15 Endopeptidase inhibitor activity CU694370.9 0.005 100% Drosophila pseudoobscura Dpse\GA24674 Complete proteome Drosophila virilis Dvir\GJ22874 Complete genome Dper\GL20353 Zinc ion binding ATCC 25017 Drosophila persimilis Francisella philomiragia Drosophila melanogaster 23 24 25 Contig515:13967 Contig8590:01715 Contig3840:01556 NADH dehydrogenase activity Defense response to bacterium 26 27 Contig4225:0563 Contig4679:38421 Homo sapiens RP11-457N9 Structural constituent of ribosome AC017078.8 0.005 100% Pan troglodytes PTB-044H21 Nucleotide binding AL954251.1 0.005 100% Arabidopsis thaliana F19K19 ATP Binding AC011808.4 5.00E-06 93% Pig DNA sequence CH242-209D19 Zinc ion binding FP085370.17 0.003 100% Mus musculus EUCOMM Integral to membrane JN962323.1 0.011 100% Pan troglodytes CH251-324M23 Fatty acid transporter activity AC197406.4 0.011 100% Homo sapiens MLLT10 Zinc ion binding NG_027818. 1 0.011 100% Macaca mulatta CH250-530I2 Cholesterol 25-hydroxylase activity AC205183.3 0.011 100% Mouse RP24-145F3 RNA binding 0.011 100% Zebrafish DKEY-184E21 Oxidoreductase activity CT030159.1 1 CT027699.1 4 0.011 100% Macrohasseltia macroterantha trnL RNA binding AY756924.1 0.011 100% Homo sapiens CTD-2303L17 Carboxy-terminal domain protein kinase AC127034.2 0.011 100% Arabidopsis thaliana BAC F19K19 Complete proteome AC011808.4 5.00E-06 93% Pig DNA sequence CH242-209D19 Zinc ion binding FP085370.17 0.003 100% Mus musculus EUCOMM Integral to membrane JN962323.1 0.011 100% 28 Contig5845:0796 Pan troglodytes CH251-324M23 Fatty acid transporter activity AC197406.4 0.011 100% Homo sapiens MLLT10 Zinc ion binding NG_027818. 1 0.011 100% Macaca mulatta CH250-530I2 Cholesterol 25-hydroxylase activity AC205183.3 0.011 100% Mouse RP24-145F3 RNA binding 0.011 100% Zebrafish DKEY-184E21 Oxidoreductase activity CT030159.1 1 CT027699.1 4 0.011 100% Macrohasseltia macroterantha trnL RNA binding AY756924.1 0.011 100% Homo sapiens CTD-2303L17 Carboxy-terminal domain protein kinase AC127034.2 0.011 100% Drosophila persimilis Dper\GL10834 Zinc ion binding XM_002015 746.1 6.00E-04 100% Oryza sativa OSJNBa0023G1 1 ATP binding AL845345.9 6.00E-04 100% Artemisia annua CAZI3763.CAZI structural constituent of ribosome EY104502.1 0 99% CCPU963 CA-responsive protein GE585313.1 2.00E-15 92% F19K19 ATP Binding AC011808.4 5.00E-06 93% CH242-209D19 Zinc ion binding FP085370.17 0.003 100% A1 AND D1 29 30 Contig10105:01419 Contig4225:0563 Cynara cardunculus Arabidopsis thaliana Pig DNA sequence 31 Contig4679:38421 Mus musculus EUCOMM Integral to membrane JN962323.1 0.011 100% Pan troglodytes CH251-324M23 Fatty acid transporter activity AC197406.4 0.011 100% Homo sapiens MLLT10 Zinc ion binding NG_027818. 1 0.011 100% Macaca mulatta CH250-530I2 Cholesterol 25-hydroxylase activity AC205183.3 0.011 100% Mouse RP24-145F3 RNA binding 0.011 100% Zebrafish DKEY-184E21 Oxidoreductase activity CT030159.1 1 CT027699.1 4 0.011 100% Macrohasseltia macroterantha trnL RNA binding AY756924.1 0.011 100% Homo sapiens CTD-2303L17 Carboxy-terminal domain protein kinase AC127034.2 0.011 100% Arabidopsis thaliana BAC F19K19 Complete proteome AC011808.4 5.00E-06 93% Pig DNA sequence CH242-209D19 Zinc ion binding FP085370.17 0.003 100% Mus musculus EUCOMM Integral to membrane JN962323.1 0.011 100% Pan troglodytes CH251-324M23 Fatty acid transporter activity AC197406.4 0.011 100% Homo sapiens MLLT10 Zinc ion binding NG_027818. 1 0.011 100% Macaca mulatta CH250-530I2 Cholesterol 25-hydroxylase activity AC205183.3 0.011 100% Mouse RP24-145F3 RNA binding CT030159.1 1 0.011 100% 32 33 34 35 36 37 38 Contig296:0295 Contig5601:0759 Contig718:01984 Contig7540:01080 Contig8220:0844 Contig209:0247 Contig249:21956 Zebrafish DKEY-184E21 Oxidoreductase activity CT027699.1 4 0.011 100% Macrohasseltia macroterantha trnL RNA binding AY756924.1 0.011 100% Homo sapiens CTD-2303L17 Carboxy-terminal domain protein kinase AC127034.2 0.011 100% Drosophila yakuba Dyak\GE12865 Complete proteome XM_002090 280.1 3.00E-04 100% Mus musculus RP23-353C21 RNA binding AC161815.4 0.001 100% Venturia canescens Vcan025 Transporter activity DQ649259.1 0.003 100% Che8 Mycobacterium phage AY129330.1 0.003 100% Gp96 Caviid herpesvirus 2 NM_214643. 2 0.007 100% Mycobacteriopahg e Strongylocentrotus purpuratus Potassium channel regulator activity Sucrose: hydrogen symporter activity FJ119030.1 3.00E-10 94% XM_002893 200.1 0.002 97% R11198 ATP Binding BT000684.1 0.002 97% RP24-74L7 RNA binding AC155164.5 0.004 100% defensin BD2 Antibiotic FR873267.1 3.00E-04 100% MTR_4g091580 PTM XM_003608 228.1 3.00E-15 92% Pinus taeda UMN_CL124 Arabidopsis lyrata SUC2 Arabidopsis thaliana Mus musculus Mytilus galloprovincialis Medicago truncatula NFIA_020680 Structural constituent of ribosome XM_001265 256.1 1.00E-13 93% ATCC 42464 Hydrolase activity CP003005.1 2.00E-16 95% Pisum sativum PUB2 Ubiquitin-protein ligase activity L81140.1 4.00E-14 92% Drosophila grimshawi Dgri\GH22070 Complete proteome 3.00E-05 100% Drosophila yakuba Dyak\GE24309 Complete proteome XM_001987 669.1 XM_002097 700.1 4.00E-04 95% Francisella sp. TX077308 Heme binding, cytochrome c CP002872.1 0.001 100% KLTH0F07854g Translation regulator activity XM_002554 498.1 0.001 100% CBS 6340 Motor Protein CU928170.1 0.001 100% Dpse\GA24674 Complete genome 0.001 100% Dvir\GJ22874 Zinc ion binding XM_002138 246.1 XM_002054 269.1 0.001 100% ATCC 25017 Chaperone CP000937.1 0.001 92% IP09721 Sequence-specific DNA binding BT029275.1 0.001 100% Eriocheir sinensis Esc29 Repressor DQ450915.1 0.001 100% Cyprinid herpesvirus FL ORF148 Membrane protein ORF148 GU815504.1 3.00E-07 97% Neosartorya fischeri Myceliophthora thermophila 39 Contig3946:2446 Lachancea thermotolerans Lachancea thermotolerans Drosophila pseudoobscura Drosophila virilis Francisella philomiragia Drosophila melanogaster 40 Contig5597:01737 Koi herpesvirus Naumovozyma dairenensis Zygosaccharomyce s rouxii Tetrapisispora blattae Kuenenia stuttgartiensis Pantoea ananatis KHV-I ATP Binding DQ177346.1 3.00E-07 100% CBS Translation regulator activity HE580273.1 1.00E-06 97% CBS732 Translation regulator activity CU928173.1 1.00E-06 97% CBS 6284 Translation regulator activity HE806321.1 1.00E-05 97% KUST_C Repressor CT573073.1 1.00E-05 97% PA13 Catalytic activity CP003085.1 2.00E-04 100% Shikimate kinase activity CP003011.1 2.00E-04 100% Thielavia terrestris NRRL 8126 Zygosaccharomyce s rouxii CBS732 Translation regulator activity CU928175.1 2.00E-04 100% Neurospora crassa B11B23 Pyridoxal phosphate binding AL669991.1 2.00E-04 100% Trypanosoma brucei gambiense DAL972 Endonuclease activity FN554971.1 6.00E-04 100% Capsicum annuum CaCM529L23 Vasodilation GU048888.1 0.002 100% C08HBa0210J04 Exonuclease AP009282.1 0.002 100% CgPTK-l ATP binding AB098199.1 4.00E-06 97% DDB_G0280785 DNA binding 4.00E-06 97% ketoacyl-CoA thiolase A Acetyl-CoA CAcyltransferase activity XM_635935. 1 XM_003329 597.1 2.00E-04 100% Solanum lycopersicum Codonosiga gracilis Dictyostelium discoideum Puccinia graminis Solanum lycopersicum LEFL1039CB09 Receptor tyrosine-protein kinase Drosophila virilis Dvir\GJ13854 Complete genome Arabidopsis thaliana IKU1 Regulation of seed growth Artemisia annua CAZI3603 Structural constituent of ribosome Lotus japonicus BP056955 Solanum tuberosum AK322583.1 6.00E-04 94% XM_002048 196.1 NM_129075. 3 6.00E-04 94% 0.002 100% EY104210.1 6.00E133 96% DNA binding GO022661.1 2.00E-08 97% POADE79 Trichodiene synthase activity CK274023.1 5.00E-05 97% Pinus pinaster PPSW21 Oxidoreductase BX000758.1 5.00E-05 95% Honey bee BH10057G22 RNA binding DB744146.1 2.00E-04 100% GE606757.1 6.00E-04 94% GE606757.1 6.00E-04 94% Cynara cardunculus Cynara cardunculus CCPX3860 CCPX3860 Serine-type endopeptidase activity Serine-type endopeptidase activity A2 AND D1 41 Contig10105:01419 Artemisia annua CAZI3763.CAZI Structural constituent of ribosome EY104502.1 0 99% Cynara cardunculus CCPU963 Heme binding, cytochrome c GE585313.1 2.00E-15 92% 42 Contig296:0295 Drosophila yakuba Dyak\GE12865 Complete proteome XM_002090 280.1 3.00E-04 100% Mus musculus RP23-353C21 RNA binding AC161815.4 0.001 100% Venturia canescens Vcan025 Transporter activity DQ649259.1 0.003 100% Che8 Mycobacterium phage AY129330.1 0.003 100% Dgri\GH22070 Complete proteome 3.00E-05 100% Drosophila yakuba Dyak\GE24309 Complete proteome XM_001987 669.1 XM_002097 700.1 4.00E-04 95% Francisella sp. TX077308 Heme binding, cytochrome c CP002872.1 0.001 100% KLTH0F07854g Translation regulator activity XM_002554 498.1 0.001 100% CBS 6340 Motor Protein CU928170.1 0.001 100% Dpse\GA24674 Complete genome 0.001 100% Dvir\GJ22874 Zinc ion binding XM_002138 246.1 XM_002054 269.1 0.001 100% ATCC 25017 Chaperone CP000937.1 0.001 92% IP09721 Sequence-specific DNA binding BT029275.1 0.001 100% Eriocheir sinensis Esc29 Repressor DQ450915.1 0.001 100% Medicago truncatula mte1-58c24 Transcription factor CR955005.2 4.00E-12 93% Mycobacteriopahg e Drosophila grimshawi 43 Contig3946:2446 Lachancea thermotolerans Lachancea thermotolerans Drosophila pseudoobscura Drosophila virilis Francisella philomiragia Drosophila melanogaster 44 Contig4155:01303 45 46 Contig451:212271 Contig5601:0759 Homo sapiens RP11-2B15 Ribonucleoprotein AC009446.1 1 1.00E-05 100% Mouse DNA RP23-278I21 Ribonucleoprotein AL596215.7 1.00E-05 100% Mus musculus KOMP Nodulation protein JN959018.1 4.00E-05 100% Homo sapiens TRAPPC9 Activator of NF-kappa-B NG_016478. 1 4.00E-05 97% Mus musculus BAC RP23118E20 Component of MHC I AC153894.8 4.00E-05 100% Homo sapiens CCBE1 Calcium ion binding 2.00E-04 100% Pig DNA CH242-140A18 Hydrolase NG_016990. 1 CU062440.4 1 2.00E-04 100% Pan troglodytes CH251-272H6 Fatty acid transporter activity AC193771.2 2.00E-04 100% Gallus gallus CH261-88N9 ATP binding AC186838.4 2.00E-04 100% Human DNA MDN1, RPL22, CASP8AP2 Ribonucleoprotein 2.00E-04 100% Homo sapiens ZMIZ1 Transcription factor AL353692.1 4 NG_028289. 1 6.00E-04 100% Macaca mulatta CH250-188A12 Receptor AC214204.5 6.00E-04 95% Rhesus Macaque CH250-352A6 Receptor AC189964.2 6.00E-04 97% Lotus japonicus LjT09E08, TM0265 DNA binding AP006391.1 6.00E-04 100% Strongylocentrotus purpuratus Gp96 Caviid herpesvirus 2 NM_214643. 2 0.007 100% 47 48 49 Contig7219:0736 Contig5987:0452 Contig5597:01737 Lotus japonicus Solanum lycopersicum Solanum lycopersicum Lycopersicon esculentum Plasmodium knowlesi Cyprinid herpesvirus LjFL2-006BC04 DNA binding AK337989.1 2.00E-10 94% LEFL1016AF02 rRNA binding AK321003.1 2.00E-13 93% AK246443.1 2.00E-13 93% BT014562.1 1.00E-10 92% FC11AF10 134016F NADH dehydrogenase activity RNA-directed DNA polymerase activity complete genome Complete genome AM910986.1 2.00E-08 94% FL ORF148 Membrane protein GU815504.1 3.00E-07 97% Geobacter sp M18, complete genome DNA damage-binding protein CP002479.1 3.00E-07 100% Koi herpesvirus KHV-I DNA polymerase DQ177346.1 3.00E-07 100% CBS Trimethyllysine dioxygenase HE580273.1 1.00E-06 97% CU928173.1 1.00E-06 97% CT573073.1 1.00E-05 97% Naumovozyma dairenensis Zygosaccharomyce s rouxii Kuenenia stuttgartiensis CBS732 KUST_C Mitogen-activated protein kinase Fragile X mental retardation protein Pantoea ananatis PA13 DNA repair protein CP003085.1 2.00E-04 100% Thielavia terrestris NRRL 8126 Transferase activity CP003011.1 2.00E-04 100% Zygosaccharomyce s rouxii CBS732 Mitogen-activated protein kinase CU928175.1 2.00E-04 100% Neurospora crassa B11B23 DNA binding AL669991.1 2.00E-04 100% Trypanosoma brucei gambiense DAL972 Nuclease activity FN554971.1 6.00E-04 100% Capsicum annuum CaCM529L23 Transfer of acetyl-CoA GU048888.1 0.002 100% C08HBa0210J04 Carboxypeptidase AP009282.1 0.002 100% CgPTK-l Tyrosine-protein kinase AB098199.1 4.00E-06 97% DDB_G0280785 DNA damage-binding 4.00E-06 97% Puccinia graminis ketoacyl-CoA thiolase A Acyltransferase XM_635935. 1 XM_003329 597.1 2.00E-04 100% Solanum lycopersicum LEFL1039CB09 Receptor tyrosine-protein kinase AK322583.1 6.00E-04 94% Drosophila virilis Dvir\GJ13854 Complete genome 6.00E-04 94% Arabidopsis thaliana IKU1 Transciptional level XM_002048 196.1 NM_129075. 3 0.002 100% Artemisia annua CAZI3603 Structural constituent of ribosome EY104210.1 6.00E133 96% Lotus japonicus BP056955 Catalytic activity GO022661.1 2.00E-08 97% EST004950 Telomerase complex DW492938.1 5.00E-05 97% POADE79 Trichodiene synthase activity CK274023.1 5.00E-05 97% PPSW21 Endonuclease activity BX000758.1 5.00E-05 95% Solanum lycopersicum Codonosiga gracilis Dictyostelium discoideum Gossypium hirsutum Solanum tuberosum Pinus pinaster 50 Honey bee BH10057G22 RNA binding DB744146.1 2.00E-04 100% Cynara cardunculus CCPX3860 Target cell lysis GE606757.1 6.00E-04 94% Complete genome EU549769.1 0 Complete genome DQ383816.1 0 99% Complete genome DQ383815.1 0 99% Complete genome HQ234669.1 0 98% ATPase HQ664599.1 0 98% Complete genome DQ898156.1 0 98% Complete genome GU456628.1 0 97% HM100310.1 0 98% HM100311.1 0 98% EU016780.1 0 98% AB240139.1 0 98% AC AND D1 Guizotia abyssinica Lactuca sativa Helianthus annuus Jacobaea vulgaris Aucuba japonica 51 Contig10461:0582 Daucus carota Anthriscus cerefolium Davidia involucrata Franklinia alatamaha Anethum graveolens Nicotiana tomentosiformis Complete genome Complete genome Complete genome Complete genome rpl23, trnI-CAU, ycf2 Complete genome Complete genome ndhB ndhB ndhB Complete genome Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Complete genome 99% Datura stramonium Solanum tuberosum Solanum bulbocastanum Buxus microphylla Coffea arabica Olea europaea Pachysandra terminalis Fosterella caulescens Japonolirion osense Doryanthes palmeri Complete genome Complete genome Complete genome Complete genome Complete genome Complete genome rps12, ndhB ndhB ndhB ndhB Pitcairnia feliciana ndhB Flagellaria indica ndhB Molineria capitulata ndhB Musa acuminata ndhB Complete genome JN654342.1 0 98% Complete genome JF772170.2 0 98% Complete genome DQ347958.1 0 98% Complete genome EF380351.1 0 98% Complete genome EF044213.1 0 98% Complete genome GU228899.2 0 97% Ribonucleoprotein AY237141.1 0 98% Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H HQ180947.1 0 97% JQ069014.1 0 97% JQ276504.1 0 97% HQ180964.1 0 97% HQ180946.1 0 97% HQ180941.1 0 97% EU017042.1 0 97% 52 Contig2126:01124 Shuttles electrons from NAD(P)H Shuttles electrons from NAD(P)H EU016718.1 0 97% HQ180959.1 0 97% Aminoacyl-tRNA synthetase JN710464.1 0 97% Os02g0594100 Transcription factor NM_001053 839.1 2.00E-07 97% Gossypium hirsutum MONCS2109 NF-kappa-B activation GQ395086.1 6.00E-07 95% Mus musculus RP24-371P8 Ribonucleoprotein AC111014.6 2.00E-06 95% Prunus persica KNOPE2 GU144519.1 8.00E-06 95% Amborella trichopoda AT_SBa0088I11 AC243606.1 8.00E-06 95% Zebrafish DKEY-12H9 ATP binding 8.00E-06 97% Mouse DNA RP23-219C18 Ribonucleoprotein AL954831.1 7 AL672141.1 2 8.00E-06 95% CH252-230M14 Cholesterol biosynthetic enzymes AC239272.5 3.00E-05 97% mth2-124i6 Transcription factor AC157982.2 7 3.00E-05 97% Y46E12BM Binding FO081797.1 3.00E-05 97% bRB-333H21 Endonuclease FO117624.1 3 1.00E-04 93% Acorus americanus ndhB Neoastelia spectabilis ndhB Asclepias macrotis LSUS Oryza sativa Chlorocebus aethiops Medicago truncatula Caenorhabditis elegans Rat DNA Sequence-specific DNA binding Cation transmembrane transporter activity 53 54 55 Contig362:0275 Contig418:22747 Contig4225:0563 Rattus norvegicus RNECO-252C13 TRANSCRIPTION FACTOR AC245774.3 1.00E-04 97% Talaromyces stipitatus UbiD/Ubi4 Covalent Attachment XM_002341 825.1 4.00E-08 100% Aspergillus niger An04c0190 Thiosulfate sulfurtransferase activity AM270080.1 4.00E-08 100% Arabidopsis thaliana UBQ4 Covalent Attachment NM_122069. 3 5.00E-07 97% Glycine max Subi-1 Covalent Attachment D16248.1 8.00E-05 94% mth2-72p17 Transcription factor CT573507.2 8.00E-04 94% F19K19 Transferase AC011808.4 5.00E-06 93% Pig DNA sequence CH242-209D19 Nucleic acid binding FP085370.17 0.003 100% Mus musculus EUCOMM MHC class II antigen JN962323.1 0.011 100% Pan troglodytes CH251-324M23 MHC II Membrane AC197406.4 0.011 100% Homo sapiens MLLT10 Zinc ion binding NG_027818. 1 0.011 100% Macaca mulatta CH250-530I2 Receptor AC205183.3 0.011 100% Vitis vinifera VV78X115886.6 Isomerase AM489072.2 0.011 100% Mouse RP24-145F3 Ribonucleoprotein 0.011 100% Zebrafish DKEY-184E21 Membrane component CT030159.1 1 CT027699.1 4 0.011 100% Medicago truncatula Arabidopsis thaliana 56 Contig2380:01787 Macrohasseltia macroterantha trnL RNA binding AY756924.1 0.011 100% Homo sapiens CTD-2303L17 Carboxy-terminal domain protein kinase AC127034.2 0.011 100% 98% Artemisia annua CAZI23588 Beta-lactam antibiotic EY080578.1 0.00E+0 0 Lactuca virosa CLVY6837 ATP binding DW168485.1 0.00E+0 0 91% Centaurea solstitialis CNSM11838 Chaperone EH765839.1 0.00E+0 0 92% 91% Taraxacum officinale CTOY4902 RNA-binding DY829234.1 0.00E+0 0 Centaurea maculosa CNMM5610 Shaping and sizing of PSM) EH732731.1 0.00E+0 0 91% Helianthus exilis CHEM3111 Receptor for the attracting Laspartate EE643217.1 0.00E+0 0 90% Silene diclinis X3 Rigid and resistant hair shaft EU521745.1 5.00E-49 90% Drosophila willistoni Dwil\GK14864 Component of nucleosome XM_002065 045.1 4.00E-05 95% Eimeria tenella AMA1 Activator protein JN032081.1 0.007 100% A1 AND DC1 57 Contig1852:02203 58 59 Contig1852:02203 Contig2157:15764 60 Leptosphaeria maculans Solanum lycopersicum Medicago truncatula JN3 lm Endonuclease FP929126.1 0.007 100% LEFL1007BC07 Wnt signaling pathway AK320272.1 0.007 100% MTR_4g113140 DNA binding XM_003609 160.1 2.00E-06 97% 95% Artemisia annua CAZI5271 Ribosomal protein EY083474.1 0.00E+0 0 Artemisia annua CAZI5271 Ribosomal protein EY107471.1 7.00E-31 100% Coffea arabica CA00-XX-FB2022-A07-JE Nucleotidyltransferase GT702383.1 1.00E-04 97% Beta-lactam antibiotic EY080578.1 0.00E+0 0 98% 91% Artemisia annua CAZI23588 61 Lactuca virosa CLVY6837 ATP binding DW168485.1 0.00E+0 0 62 Centaurea solstitialis CNSM11838 Chaperone EH765839.1 0.00E+0 0 92% 63 Taraxacum officinale CTOY4902 RNA-binding DY829234.1 0.00E+0 0 91% 64 Centaurea maculosa CNMM5610 shaping and sizing of PSM) EH732731.1 0.00E+0 91% Contig2380:01787 0 65 66 Contig5601:0759 Helianthus exilis CHEM3111 Receptor for the attracting Laspartate Strongylocentrotus purpuratus Gp96 Caviid herpesvirus 2 Zebrafish 67 Contig7708:01368 68 Contig1852:02203 69 Contig5184:30251 70 Contig1514:18187 CH73-166J22 EE643217.1 Endopeptidase inhibitor activity NM_214643. 2 CU856173.1 9 0.00E+0 0 90% 0.007 100% 2.00E-17 92% Pongo abelii CH276-98H12 Serine esterase AC187773.2 2.00E-17 100% Zebrafish SLC8A2 Calcium: sodium antiporter activity BX890602.8 8.00E-17 93% Rattus norvegicus BN/SsNHsd/MC W Electron carrier activity AC133262.3 3.00E-16 92% Mus musculus RP24-323H7 Nucleic acid binding AC102594.1 5 1.00E-15 91% Macaca mulatta DRB Immune system AM910410.1 1.00E-15 90% Atelerix albiventris LBNL4-89B6 Positive regulator of the HHSP AC144402.2 4.00E-15 91% Silene diclinis X3 Keratin-associated protein EU521745.1 5.00E-49 90% Mouse DNA RP23-407I21 Isomerase, topoisomerase 9.00E-05 100% Canis familiaris XX-297J22 Probable collagen protein 0.001 97% Giardia lamblia GL50803_16310 Ribonucleoprotein AL603706.1 3 AC188148.1 6 XM_001703 998.1 9.00E-09 91% 71 Contig851:01951 Senecio vernalis eifsv1 Protein biosynthesis AJ238624.1 3.00E173 Brassica oleracea BoCAR44 Putative uncharacterized protein DQ267199.1 6.00E-64 90% Silene diclinis X3 Keratin-associated protein EU521745.1 5.00E-49 90% Artemisia annua CAZI5271 Organization of the cytomatrix EY083474.1 0.00E+0 0 95% Coffea arabica CA00-XX-FB2022-A07-JE RNA binding GT702383.1 1.00E-04 97% Gp96 Caviid herpesvirus 2 0.007 100% Dper\GL10834 Zinc ion binding NM_214643. 2 XM_002015 746.1 6.00E-04 100% Oryza sativa OSJNBa0023G1 1 zinc ion binding AL845345.9 6.00E-04 100% Zebrafish CH73-166J22 Nucleic acid binding CU856173.1 9 2.00E-17 92% Pongo abelii CH276-98H12 Serine esterase AC187773.2 2.00E-17 100% Zebrafish SLC8A2 Calcium: sodium antiporter activity BX890602.8 8.00E-17 93% Rattus norvegicus BN/SsNHsd/MC W Electron carrier activity AC133262.3 3.00E-16 92% 90% A2 AND DC1 72 73 74 Contig161:0979 Contig1852:02203 Contig2157:15764 75 76 77 78 79 Contig5601:0759 Contig584:0908 Contig7708:01368 Strongylocentrotus purpuratus Drosophila persimilis Mus musculus RP24-323H7 Zinc ion binding AC102594.1 5 1.00E-15 91% Macaca mulatta DRB MHC II BINDING AM910410.1 1.00E-15 90% Atelerix albiventris LBNL4-89B6 Positive regulator of HHP AC144402.2 4.00E-15 91% 80 Contig851:01951 Senecio vernalis eifsv1 protein biosynthesis AJ238624.1 3.00E173 90% 81 Contig4155:01303 Medicago truncatula mte1-58c24 Acyl-CoA thioesterases CR955005.2 4.00E-12 93% Homo sapiens RP11-2B15 Ribonucleoprotein AC009446.1 1 1.00E-05 100% Mouse DNA RP23-278I21 Ribonucleoprotein AL596215.7 1.00E-05 100% Mus musculus KOMP Nodulation protein JN959018.1 4.00E-05 100% Homo sapiens TRAPPC9 Activator of NF-kappa-B NG_016478. 1 4.00E-05 97% Mus musculus BAC RP23118E20 COMPONENT OF MHC I AC153894.8 4.00E-05 100% Homo sapiens CCBE1 Calcium ion binding 2.00E-04 100% Pig DNA CH242-140A18 Hydrolase NG_016990. 1 CU062440.4 1 2.00E-04 100% Pan troglodytes CH251-272H6 MHC II Membrane AC193771.2 2.00E-04 100% Gallus gallus CH261-88N9 ATP binding AC186838.4 2.00E-04 100% 82 Contig451:212271 83 84 Human DNA MDN1, RPL22, CASP8AP2 Ribonucleoprotein 2.00E-04 100% TRANSCRIPTION FACTOR AL353692.1 4 NG_028289. 1 Homo sapiens ZMIZ1 6.00E-04 100% Macaca mulatta CH250-188A12 Receptor AC214204.5 6.00E-04 95% Rhesus Macaque CH250-352A6 Receptor AC189964.2 6.00E-04 97% Lotus japonicus LjT09E08, TM0265 DNA binding AP006391.1 6.00E-04 100% Drosophila grimshawi Dgri\GH22070 Complete proteome 3.00E-05 100% Drosophila yakuba Dyak\GE24309 Complete proteome XM_001987 669.1 XM_002097 700.1 4.00E-04 95% Francisella sp. TX077308 Heme binding, cytochrome c CP002872.1 0.001 100% KLTH0F07854g Translation regulator activity XM_002554 498.1 0.001 100% CBS 6340 Motor Protein CU928170.1 0.001 100% Dpse\GA24674 Complete genome 0.001 100% Drosophila virilis Dvir\GJ22874 Zinc ion binding XM_002138 246.1 XM_002054 269.1 0.001 100% Francisella philomiragia ATCC 25017 Chaperone CP000937.1 0.001 92% Drosophila IP09721 Sequence-specific DNA BT029275.1 0.001 100% AC AND DC Contig3946:2446 Lachancea thermotolerans Lachancea thermotolerans Drosophila pseudoobscura binding melanogaster Eriocheir sinensis Esc29 Repressor DQ450915.1 0.001 100% 85 Contig418:22747 Medicago truncatula mth2-72p17 Involved in biological aging CT573507.2 8.00E-04 94% 86 Contig8369:111513 Artemisia annua CTOX5296.b1_ P04 Thymocyte marker EY115636.1 1.00E163 95% 87 Contig8590:01715 Solanum lycopersicum hba-17c15 map 1 Involved in oxygen transport AC246154.3 5.00E-96 90% Artemisia annua CAZI22903 Beta-lactam antibiotic EY096790.1 0 99% 88 Contig8879:0594 Chrysanthemum x morifolium LMS Ribonucleoprotein DK939321.1 2.00E-95 98% Mouse DNA Rps2 Disease resistance (R) protein AL772210.8 6.00E-06 97% Rat DNA bRB-233C6 Endonuclease FO181541.1 1 7.00E-05 100% Mizuhopecten yessoensis HLJX-141 Chaperone FJ262385.1 7.00E-05 100% Zebrafish DKEY-153P21 Oxidoreductase activity CU633478.1 0 7.00E-05 100% Rattus norvegicus SHR-Akr Endopeptidase inhibitor activity AC242886.6 3.00E-04 97% Pan troglodytes CH251-381H2 Fatty acid transporter activity AC192852.3 3.00E-04 100% Mus musculus RP24-127A22 Ribonucleoprotein AC123833.3 3.00E-04 100% Oryza sativa OSJNBa0033D2 4 ATP binding AP005439.3 3.00E-04 100% 89 Contig8881:0761 Tetrapisispora phaffii Gossypium herbaceum CBS 4417 Motor Protein HE612859.1 0.001 100% NBRI_M189 GTPase activity JF495680.1 0.001 100% Homo sapiens FXYD6 Ion channel activity NG_013071. 1 0.001 100% Villosa fabalis VfaB244 Hydrolase activity GQ258763.1 0.001 100% Candida dubliniensis CD36_06900 Protein-degradation mediator XM_002417 302.1 0.001 100% Felis catus FCAB-91H12 Carbonate dehydratase activity AC234574.1 0.001 100% Dictyostelium discoideum DDB_G0282573 DNA binding XM_635042. 1 0.001 100% Pan troglodytes RP43-148A6 Fatty acid transporter activity AC146001.2 0.001 100% Zebrafish CH211-218I18 Nucleic acid binding AF324786.1 0.001 100% Homo sapiens RP11-728F11 Ribonucleoprotein AP000757.4 0.001 100% Weeksella virosa DSM 16922 Nucleotidyltransferase CP002455.1 0.003 100% Pongo abelii CH276-480A19 DNA topoisomerase activity AC206881.3 0.003 100% Thais clavigera Tcl128 ATP synthase FJ550296.1 Mus musculus RP24-267D17 Ribonucleoprotein AC123055.3 1.00E-05 100% Homo sapiens RP11-513N24 Ribonucleoprotein AC022164.8 1.00E-05 100% 1.00E-05 90 Contig8890:0844 100% 91 92 Contig2380:01787 Contig4225:0563 Homo sapiens PRKCB Transferase, ATP Binding NG_029003. 1 1.00E-05 100% Gossypium hirsutum NBRI_Gh_J014 GTPase activity HQ524507.1 0.002 97% 98% Artemisia annua CAZI23588 Beta-lactam antibiotic EY080578.1 0.00E+0 0 Lactuca virosa CLVY6837 ATP binding DW168485.1 0.00E+0 0 91% Centaurea solstitialis CNSM11838 Chaperone EH765839.1 0.00E+0 0 92% 91% Taraxacum officinale CTOY4902 RNA-binding DY829234.1 0.00E+0 0 Centaurea maculosa CNMM5610 shaping and sizing of PSM) EH732731.1 0.00E+0 0 91% Helianthus exilis CHEM3111 Receptor for the attracting Laspartate EE643217.1 0.00E+0 0 90% Arabidopsis thaliana F19K19 ATP Binding AC011808.4 5.00E-06 93% Pig DNA sequence CH242-209D19 Zinc ion binding FP085370.17 0.003 100% Mus musculus EUCOMM Integral to membrane JN962323.1 0.011 100% Pan troglodytes CH251-324M23 Fatty acid transporter activity AC197406.4 0.011 100% Homo sapiens MLLT10 Zinc ion binding NG_027818. 1 0.011 100% Macaca mulatta CH250-530I2 Cholesterol 25-hydroxylase activity AC205183.3 0.011 100% Mouse RP24-145F3 RNA binding 0.011 100% Zebrafish DKEY-184E21 Oxidoreductase activity CT030159.1 1 CT027699.1 4 0.011 100% Macrohasseltia macroterantha trnL RNA binding AY756924.1 0.011 100% Homo sapiens CTD-2303L17 Carboxy-terminal domain protein kinase AC127034.2 0.011 100% Artemisia annua CAZI23588 Beta-lactam antibiotic EY080578.1 0.00E+0 0 98% 91% D1 AND DC 93 Contig2380:01787 Lactuca virosa CLVY6837 ATP binding DW168485.1 0.00E+0 0 Centaurea solstitialis CNSM11838 Chaperone EH765839.1 0.00E+0 0 92% 91% 91% Taraxacum officinale CTOY4902 RNA-binding DY829234.1 0.00E+0 0 Centaurea maculosa CNMM5610 shaping and sizing of PSM) EH732731.1 0.00E+0 0 94 Contig362:0275 Helianthus exilis CHEM3111 Receptor for the attracting Laspartate EE643217.1 0.00E+0 0 90% Talaromyces stipitatus UbiD/Ubi4 Covalent Attachment XM_002341 825.1 4.00E-08 100% Aspergillus niger An04c0190 Transferase AM270080.1 4.00E-08 100% Arabidopsis thaliana UBQ4 Covalent Attachment NM_122069. 3 5.00E-07 97% Glycine max Subi-1 Covalent Attachment D16248.1 8.00E-05 94% XM_001987 669.1 XM_002097 700.1 3.00E-05 100% 4.00E-04 95% Drosophila grimshawi 95 Contig3946:2446 Dgri\GH22070 Complete proteome Drosophila yakuba Dyak\GE24309 Complete proteome Francisella sp. TX077308 Heme binding, cytochrome c CP002872.1 0.001 100% KLTH0F07854g Translation regulator activity XM_002554 498.1 0.001 100% CBS 6340 Motor Protein CU928170.1 0.001 100% Dpse\GA24674 Complete genome 0.001 100% Drosophila virilis Dvir\GJ22874 Zinc ion binding XM_002138 246.1 XM_002054 269.1 0.001 100% Francisella philomiragia ATCC 25017 Chaperone CP000937.1 0.001 92% Lachancea thermotolerans Lachancea thermotolerans Drosophila pseudoobscura 96 97 Contig7708:01368 Contig848:0283 Drosophila melanogaster IP09721 Sequence-specific DNA binding BT029275.1 0.001 100% Eriocheir sinensis Esc29 Repressor DQ450915.1 0.001 100% Zebrafish CH73-166J22 Nucleic acid binding CU856173.1 9 2.00E-17 92% Pongo abelii CH276-98H12 Serine esterase AC187773.2 2.00E-17 100% Zebrafish SLC8A2 Calcium: sodium antiporter activity BX890602.8 8.00E-17 93% Rattus norvegicus BN/SsNHsd/MC W Electron carrier activity AC133262.3 3.00E-16 92% Mus musculus RP24-323H7 Zinc ion binding AC102594.1 5 1.00E-15 91% Macaca mulatta DRB MHC II BINDING AM910410.1 1.00E-15 90% Atelerix albiventris LBNL4-89B6 Positive regulator of HHP AC144402.2 4.00E-15 91% Nicotiana tabacum H-N_BC2M41322 DNA binding DQ460104.1 4.00E-14 91% Solanum lycopersicum C02HBa0091J18 Transferase activity AC215381.1 9.00E-11 90% Likewise results of BLAST of different contigs also showed many enzymes that have established role in A.annua and A. dubia and were common to various other species. Since the time of divergence of two species, a gene might have changed its function it played in the ancestor, however, sequence still has certain homology with its evolutionary related species. BLAST of these transcriptomic sequences generated E-value that gives an idea of relatedness of sequences. If an E value is below 0.01, then the match is significant. Nonetheless, if the E value is higher, the match might still be significant especially for a short sequence. In that case, results have to be evaluated in the context of experiment. This computational and predictive analysis of features, common amongst different organisms, gives a reliable and significant amount of data. These results can have many implications and can be extrapolated to understand different unknown aspects for a newly sequenced genome. There are many organisms that show 100% sequence conservation with contigs generated through alignment of different samples of A. annua and A.dubia. The genes of these organisms are well characterized and can be helpful in predicting the molecular functions of genes of these Artemisia species. 6.4.4: Contigs of A.annua and A.dubia producing 80-85% homology BLAST of Various contigs which were obtained by aligning different transcriptomic sequences of A.annua and A.dubia, was performed and sequence homology presented interesting results. These contigs showed homology with different genome sequences ranging from 60% to 90% but only those Contigs were analyzed that showed 90% homology or more than it (Table 6.2). However there are many contigs which showed 80% and more homology produced valuable results. Contigs showing 80% homology can be seen in Appendix VI. 6.4.5: Contigs of A.annua and A.dubia do not produce homology The reason why so many genes did not show homology is not altogether surprising using the nr database. If EST database is used many Expressed sequence Tags are available for Artemisia. The fact is we still do not know what a lot of genes do and hence gene functions of many newly sequenced genes of Artemisia can also not be predicted. Out of 500 contigs, 244 contigs did not show any homology. However, if any gene sequence of A. annua and A. dubia does not match with gene sequence in any data base it is justified if the sequence has completely been changed by insertional and deletional mutagenesis, during the slow process of evolution. Fast sequence divergence can cause a sequence to completely change from what was handed down by the common ancestor. Such gene sequences lose homology. 6.5: Conclusion In this part, the transcriptome of Artemisia annua and Artemisia dubia were sequenced for the first time. Gene contigs were generated by alignment of the sequences derived from a compilation of all the sequences obtained from a species. The different individual samples from both Artemisia annua and Artemisia dubia were then compared against this reference gene list to evaluate if any genes in the genome of the transformed plants are up or down regulated compared to untransformed plants.  Genes of transgenic lines of Artemisia annua and Artemisia dubia were found to be highly up regulated as compared to control plants.  Transgenic lines carrying two copy numbers were found up regulated compared to transgenic lines having one copy number.  Transformed and untransformed plants of A.annua showed up regulation compared with transformed and untransformed plants of A.dubia.  BLAST of different contigs also showed many enzymes that have established role in A.annua and A. dubia and were common to various other species.  These results can have many implications and can be extrapolated to understand different unknown aspects for a newly sequenced genome. Chapter No. 7 Discussion DISCUSSION Throughout the history plants have been of great importance to medicine. Plants with medicinal properties have been utilized successfully in the treatment of ailments of varying degrees of severity. The compounds that are responsible for medicinal property of the drugs are usually secondary metabolites (Vormfelde, and Paser, 2000). The use of the medicinal plants for curing disease has been documented in history of all civilizations (Street, 1977). One such genus is Artemisia; different species of Artemisia are famous for their medicinal properties like Artemisia dubia and Artemisia annua. Artemisia species produce a very effective antimalarial compound artemisinin, which also showed activities against cancer cells, schistosomiasis, and certain viruses, i.e., human cytomegalovirus, hepatitis B and C virus, and bovine viral diarrhea virus. Interestingly, the antimicrobial bioactivity of artemisinin is not just against Plasmodium but it seems to be even broader and also includes the inhibition of other protozoan such as Leishmania, Trypanosoma, and Toxoplasma gondii, as well as some trematodes, fungi, yeast, and bacteria. The analysis of its complete profile of pharmacological activities, as well as the elucidation of molecular modes of action and the performance of clinical trials, will further elucidate the full potential of this versatile weapon from nature against diseases (Efferth, 2009). Since the discovery and elucidation of chemical structure of artemisinin, several teams reported different pathways for the synthesis of artemisinin, but all of them had many steps and a low overall yield. None of these complex synthetic pathways provide a feasible method for the large-scale production of artemisinin. At the moment, its extraction from Artemisia species plants remains the only commercial source for the drug (Avery et al., 2003) and it its concentration varies from 0.01 to 1.5 % DW (Christen and Veuthey, 2001; Ferreira et al., 2005). Over the past decade, novel features of plant cells transformed with the rol A, B and C genes have been revealed, such as increased production of secondary metabolites. The rol A gene has emerged as a stimulator of growth and secondary metabolism (Altvorst et al., 1992; Schmulling et al., 1993). The rol B protein on the other hand, has been shown to have a tyrosine phosphatase activity and therefore a possible role in the auxin signal transduction pathway (Filippini et al., 1996). Estruch et al. (1999) have demonstrated that rol C can be involved in the release of active cytokinins from their inactive glucosides due to its cytokinin glucosidase activity and therefore involved in activation of secondary metabolism. Collectively, these genes play a major role in the pathway that leads to high levels of secondary metabolites. Production of transgenic plants seems to be the most appropriate choice to improve production of secondary metabolites of any plant. For this purpose Agrobacterium mediated transformation method is being used for several medicinal plants including Artemisia species (Nin et al., 1996; Nin et al., 1997; Vergauwe et al., 1998). The demand of artemisinin is increasing due to its multifunctionality e.g. overall demand of artemisinin and its precursors for treatment of malaria has increased from 22,000 treatment courses in 2001 to an estimated 200 million treatment courses in 2008 (WHO, 2008 update). Therefore, the aim of this project was to enhance and compare artemisinin production in the plants of Artemisia dubia and Artemisia annua transformed with rol genes by using Agrobacterium tumefaciens and Agrobacterium rhizogenes with the non transformed plants. This report also covers the effect of different factors, such as type of explants, sterilization period and cocultivation period, on transformation efficiency and analysis of anti-cancerous activities on cancer cell lines and their comparison in transformed and nontransformed plants. In addition to it, this report also evaluates the metabolic pathways through which these rol genes enhance the production of artemisinin and to analyse its anticancerous activities. This report also explains transcriptome analysis of transformed and untransformed plants of Artemisia annua and Artemisia dubia. 7.1: Agrobacterium tumefaciens mediated transformation of Artemisia dubia and Artemisia annua with rol genes Agrbacterium mediated transformation is an effective and widely used approach to introduce foreign DNA into dicotyledons plants. For different plant species different gene transfer protocol are applicable. Various researchers have reported transformation of Artemisia species with Agrobacterium but mostly they transformed Artemisia annua. (Biswajit et al., 2000; Vergauwe et al., 1998; Vergauwe et al., 2002a, Vergauwe et al., 2005). Little work has been done on Artemisia dubia (Jun-Li et al., 2005, Kaul et al., 1976; Zafar et al., 1990; Nin et al., 1996; 1997; Rizwana et al., 2002, Altvorst et al., 1992, Charles et al., 1990, Davioud et al., 1988, Giri and Narasu, 2000). Before the transformation protocol established by Vergauwe et al, (1996b) no successful procedure had been developed to regenerate Artemisia plants in short time. Mannan et al, (2008) transformed the Artemisia dubia with Agrobacterium rhizogenes and obtained transformed regenerated plants and in 1998, Vergauwe studied some more important factors that can affect the transformation efficiency. The results from these studies were combined to produce an enhanced protocol for the transformation with Agrobacterium tumefacienes of Artemisia dubia L. The protocol includes incubation of the explants for five minutes in Agrobacterium solution, containing 50mg/L kanamycin, having a pH of 5.8 and an optical density of 1 at 560nm. It was followed by a three days cocultivation period on MS medium supplied with 200 µM acetosyringone and 50mg/l kanamycin. Explants were then transferred to selective regeneration medium containing 0.1 mg/L BAP, 50mg/l cefotaxime and 20 mg/L kanamycin. Rooting was obtained on half MS medium containing 0.025mg/L NAA. A large number of literatures around the globe has come in print on the transformation of Artemisia annua, but no work has been done on transformation by Agrobacterium in Artemisia annua growing naturally in Pakistan. This report explains an enhanced and efficient protocol for the transformation with Agrobacterium tumefacienes and Agrobacterium rhizogenes of Artemisia annua L. The Agrobacterium tumefacienes mediated transformation protocol includes incubation of the explants for five minutes in Agrobacterium solution, containing 50mg/L kanamycin, having a pH of 5.8 and an optical density of 1 at 560nm. Three days cocultivation period was given on MS medium supplied with 200 µM acetosyringone and 50mg/l kanamycin. Explants were then transferred to selective regeneration medium containing 1 mg/L BAP, 50mg/l cefotaxime and 15 mg/L kanamycin. Rooting was obtained on MS medium containing 0.1mg/L NAA. During Agrobacterium tumefaciens mediated transformation, proper sterilization procedure is required for in vitro germination of seeds. Mercuric chloride was used as a sterilizing agent. Sterilizing agents play important role in the germination of seeds, long duration had an inhibitory effect on seed germination and the seeds became dead while in short time treatment the percentage germination of seeds increased with minimum germination period. Two minutes treatment was found to be the most suitable duration for the sterilization of seeds (Fig. 2.2). Explants from in vitro grown seedlings of 20-days old seedlings were used for the transformation procedure. Explants were precultured on MS medium for three days. Co-cultivation period affects the transformation efficiency. MS medium (Murashige and Sakoog, 1962) containing 200 µM acetosyringone was used as cocultivation medium and after cocultivation, explants were placed on cocultivation medium for 2-days. Schmid et al. (2005) showed that this treatment increases the transformation efficiency up to 20%. It seems that after cocultivation of explants with Agrobacterium, a certain period of time prior to the shifting of explants to the selection medium positively influences the transformation efficiency. It probably allows the T-DNA transfer, integration, transcription and sufficient enzyme production leading to the expression of kanamycin resistant phenotypes. In order to select the suitable types of explants for regeneration, different type of explants were regenerated on media in order to check their regeneration efficiencies. Leaf and stem explants were used on MS medium containing 0.1mg/l BAP, 20mg/l kanamycin and 500mg/l cefotaxime. Maximum regeneration efficiency was observed by stem explants on this medium. So this medium was used as selection medium after the cocultivation of explants. After the infection with Agrobacterium, explants were placed on cocultivation medium for 2-days. Then these explants were shifted to selection medium. In selection medium 50mg/l kanamycin was used for the selection of transformed explants as used by Baroncelli et al. (1992), Jorsobe et al. (2003) and Vatsya et al. (2002). Explants infected with Agrobacterium tumefaciens strain LBA 4404 with plasmid RT99 and harbouring rol ABC genes. Among the two types of explants used, stem explants showed maximum transformation efficiency on this SM (MS medium containing 0.1mg/l BAP, 50mg/l kanamycin and 500mg/l cefotaxime). Leaf explants were not found to be much efficient in generating transgenic shoots and most of these died on the selection medium. Thus it appears that leaf explants are not much worth carrying out transformation experiments. This increased transformation efficiency of stem explants as compared to leaf explants has also been reported by Skoog and Miller (1997), Robertson and Earle (2000), Murata (1998), Gamborg et al. (1968) and Bidney et al. (1991). For the selection of transformed explants 50mg/L kanamycin and 500mg/l cefotaxime were used in the selection media. Various reports (Davioud et al., 1988, Maurel et al., 1991, Morgan et al., 2004) have shown that 50mg/l kanamycin is sufficient for the selection of transformed plants. The control of Agrobacterium growth at 500mg/l of cefotaxime has been shown by Vatsya et al. (2002), Jorsobe et al. (2003) and Bettini et al. (2003). For determination of best explants for shoot regeneration, different types of explants were tried. Considering the different kinds of explants, the use of stem explants seems best for the formation of shoots as well as production of Artemisia dubia and Artemisia annua plants. Shoot regeneration response observed on MS medium containing BAP 0.1mg/L, 50mg/l Kanamycin and 500mg/l cefotaxime with stem explants was 80-100%. Shoot regeneration response of leaf explants was not good as the leaves mostly formed calli on this medium. It is in accordance with Schmid et al, (2005) who worked on A. annua and reported that within three weeks all stem explants of A. annua developed shoots. However the shoot regeneration continued further and one explants led to the formation of clusters of several shoots, so that the number of regenerated shoots was higher than the original number of stem explants (Schmid et al., 2005). Nair et al, (2001) reported the regeneration of Artemisia plants starting from leaf explants by using a hormone combination of the auxin NAA (0.05) mg/L and the cytokinin BAP (0.1 to 0.2 mg/L). In about 60 days distinct shoots were obtained in this procedure, but the rooting of the shoots took 3 months, suggesting that an optimum concentration of hormone in the regeneration medium highly affects the time of regeneration of explants. The BAP (0.1mg/L) was found to be best for regeneration of shoots. Once the transformed shoots were obtained, they were rooted on rooting medium. In case of Artemisia dubia it occurred by using 0.025mg/L NAA in the rooting medium and in case of Artemisia annua it occurred by using 0.1mg/L NAA in the rooting medium, this is in contrary to Vergauwe et al., (1998), who while working on the Agrobacterium tumefaciens-mediated transformation of A. annua reported high efficiencies of root induction of shoots grown on non-selective medium. Their study suggested that cocultivation of the explants with Agrobacterium tumefaciens did not affect rooting of regenerated shoots. Whereas we have faced great difficulty as both the transformed and untransformed shoots were not giving roots on the proposed medium of i.e. 0.1mg/L NAA. However, rooting of Artemisia annua and Artemisia dubia is in accordance to the results of Jun-Li et al., (2005) and Nair et al., (2001) who were working with Agrobacterium tumefacienes mediated transformation of Artemisia annua and found rooting of Artemisia annua on 0.05-2.0mg/L NAA and 0.1mg/L NAA. We got rooting within 1-2 weeks in transformed and untransformed shooted explants by culturing them on rooting medium containing 0.025mg/L in case of Artemisia dubia and 0.1mg/L NAA in case of Artemisia annua. According to our results the number of roots developed on transformed shoots was restricted whereas Nin et al., (2004), while working on transformed Artemisia annua in vitro grown plants exhibited a more developed rooting system and roots appeared on the stem too. These differences could be due to the difference in the specie or more pronouncedly it is because they had used Agrobacterium rhizogenes for infection and as we know that it is a natural plant pathogen responsible for adventitious root formation at the site of infection (Hooykaas, 2004). While we have used Agrobacterium tumefaciens, which is a soil bacterium and causes crown gall disease (Brown, 2001). Other factor studied that could affect the transformation efficiency was cocultivation period. Time of incubation of explants with Agrobacterium inoculums is an important factor affecting the transformation efficiency. An incubation time of five minutes was found to be best giving 70% transformation efficiency. Increase in incubation time led to decrease in transformation efficiency because of over growth of Agrobacterium that hinders the growth of the transformants and hence has a negative effect on transformation efficiency. Transformation efficiency is also influenced by cocultivation time. Explants were cocultivated for two or three days. The highest formation of transgenic plants was generally obtained after two days of cocultivation. All the leaf and stem explants showed positive gene expression. It is in agreement with the results of Vergauwe et al. (1998) as they did cocultivation of Artemisia annua for 48 to 60 hours. According to their findings longer cocultivation periods did not help the transformation efficiency because of over growth of Agrobacterium as mentioned earlier. For successful Agrobacterium transformation, elimination of bacteria from culture soon after transformation is necessary. This is achieved by the addition of antibiotics into the culture medium. Antibiotics, which are commonly used to eliminate Agrobacterium tumefaciens, have been shown to influence the morphogenesis either positively or negatively (Jun-Li et al., 2005). Different doses of cefotaxime were used to optimize the standard dose which can control the overgrowth of Agrobacterium and also do not inhibit the regeneration. Cefotaxime at a concentration of 500mg/L controlled overgrowth of bacterium while doses higher than 500mg/L resulted in necrosis and hence no regeneration was observed. Even though this concentration of cefotaxime retarded the callus formation, regenerated plants were able to grow successfully. This is an agreement with Vergauwe et al, (1996b), who used cefotaxime as decontaminating antibiotic and reported that it results in retardation of callus formation and inhibition of the shoot inducing capacity in Artemisia annua. We have reported some morphological differences that existed among some transgenic and non-transgenic plants. Increased plant height, broad leaves, branched and hard stem are some of the features observed among transgenic plants (Table: 2.2). Such features were not found in control plants and hence seem to be associated with rol genes as increased plant height and broad leaves have been reported earlier by Mercke et al. (2000) and Wallaart et al. (2001). These morphological changes could be because of the hormonal imbalance caused by the rol genes as the rol genes have been reported to cause functional imbalance in the auxin/cytokinin ratio in favour of cytokinins (Schmulling et al., 1993). Regenerated plantlets were potted in the greenhouse, and their survival rate was 98%, which is in accordance with the findings of Fazal et al, (1997). They acclimatized the Artemisia absinthium and 95% of the plantlets were successfully transplanted to soil and continued to grow in the field. 7.2: Agrobacterium rhizogenes mediated transformation of Artemisia dubia and Artemisia annua with rol genes Hairy roots produced as a result of genetic transformation by A. rhizogenes, often grow as fast as or faster than plant cell cultures (Flores and Filner, 1985). The neoplastic roots are characterized by a high growth rate and are able to synthesize some secondary metabolites (Flores and Filner, 1985). The greatest advantage of hairy roots is that hairy root cultures often exhibit about the same or greater biosynthetic capacity for secondary metabolite production compared to their mother plants (Banerjee et al., 1998). Even in cases where secondary metabolites accumulate only in the aerial part of an intact plant, hairy root cultures have been shown to accumulate the metabolites. For example, lawsone normally accumulates only in the aerial part of the plant, but their hairy roots can produce lawsone (Bakkali et al., 1997). Hairy roots many provide the mean to produce these commercially valuable products on a large scale (Flores and Curtis, 1992; Wilson, 1997). Little work has been published about the transformation of A.dubia and A.annua with Agrobacterium rhizogenes growing naturally in Pakistan except one report by Mannan et al., (2008) transformed the A.dubia with Agrobacterium rhizogenes and generated hairy roots. The results from these studies were combined to produce an enhanced protocol for the transformation with Agrobacterium rhizogenes of Artemisia dubia WALL. and Artemisia annua L. The protocol includes incubation of twenty eight days old sterile plantlets of A. annua and Artemisia dubia. A single colony of A. rhizogenes, 24 hours incubated Agrobacterium cultures, was picked with sterile scalpel which lightly incised into the stem portion of A. annua. In each plantlet, 2-3 incisions were made on different places along the length of stem, by Agrobacterium rinsed scalpel. The infected plantlets were incubated again in growth room having 24 μM m-2 sec-1 cool white fluorescent light and 25ºC with 16/8 light period. After seven days of infection, hairy roots started emerging from Agrobacterium infected stem portions of A. annua and Artemisia dubia. In the present investigation, after transformation the hairy roots from both Artemisia species and roots from nontransformed plants were shifted to solid B5 medium for further proliferation, the transformed roots grow faster as compared to non transformed roots. Norma and Ana Maria (1995) also reported that hairy root culture can significantly increase 2.5-3 fold the secondary metabolites in comparison with control and also reported that transformed and non-transformed roots exhibit a strong tendency towards proliferation in liquid B5 medium. For transformation with Agrobacterium rhizogenes, small stem (2-4cm) portions of Artemisia dubia and Artemisia annua having a bud were transferred to MS medium (Murashige and Skoog, 1962) after sterilization. New plants appeared from bud after one month and hairy roots appeared in six days after infection with Agrobacterium rhizogenes strain LBA9402 (Fig. 2.11), while Agrobacterium rhizogenes strain LBA8196 did not produce any hairy roots. Mannan et al. (2008) also reported the production of hairy roots in Artemisia dubia and Artemisia indica with Agrobacterium rhizogenes strain LBA9402 and LBA8196. Nin et al (1997) also infected different genotypes of Artemisia absinthium shoots with two different Agrobacterium rhizogenes strain LBA1855 and LBA9402 and reported production of hairy roots with strain LBA9402 and LBA1855. Similarly, Giri and Narasu (2001) transformed Artemisia annua with different Agrobacterium rhizogenes strains such as A4, K599, LBA9402, 9365, 9340 and reported variable virulency among these strains for induction of hairy roots with the best transformation response obtained with LBA9402. These reports are in agreement with our results that LBA9402 is more effective in generating Artemisia dubia and Artemisia annua hairy roots as compared to LBA8196. Even though LBA8196 did not produce any transformed roots in Artemisia dubia and Artemisia annua yet the possibility of this strain to produce transformants in other species of Artemisia cannot be excluded. The transformation of Artemisia dubia and Artemisia annua was confirmed by PCR and Southern blot analysis. All the transformants tested gave positive PCR results for the rol A, B and C genes. Southern blots also detected the copy number in independent transgenic lines with the single T-DNA insertion being the most frequent. Han et al (2005) and Zhang et al (2009) also reported the optimization of Agrobacterium transformation systems in Artemisia annua and confirmed the integration of the target genes by Southern blot analysis and found double number in 90% plants. Waleerat et al (2010) and Hobbs et al (1993) also described the transformation of Artemisia annua with Agrobacterium tumefacienes and confirmed integration of target genes by southern blot analysis and found double number only in one plant out of 7 plants. Nin et al (1997) explained the transformation of Artemisia absinthium with Agrobacterium rhizogenes and confirmed the transformation through Southern blot analysis and did not find double copy number in any of transformed plant. 7.3: Analysis of artemisinin and its derivatives Extraction of artemisinin, artemether, arteether, artesunate and dihydroartemisinin with Ethyl acetate (Ferreira and Janick, 2002), acetonitrile (Elsohly et al., 1987), chloroform (Woerdenbag et al., 1991; Mary et al., 1994) and petroleum ether (Klayman et al., 1984) has been reported with the extraction time ranging from a few minutes (Duke et al., 1994) to several hours (Charles et al., 1990). Charles et al. (1990) reported rapid extraction and 48 hours long extraction methods of artemisinin and its derivatives and found no significant difference. In the present project, extraction of artemisinin from plant material was carried out using Ethyl acetate and hexane in the ratio of 95:5 and acetonitrile 100% as it is most widely used laboratory techniques today (Van Quekelberghe et al., 2008; Yan Liu et al., 2009 ; Alexei et al., 2010). It is worth mentioning here that some alteration in artemisinin and its derivatives concentration in previously reported work may be possible due to different methods of extraction. Mingfu et al. (2005) used 50 ml methanol to extract artemisinin from 2 g of A. annua leaves powder followed by Sonication for 45 minutes and centrifugation at 12000 rpm for 3 minutes resulting in 0.03 to 0.71 % artemisinin. Elsohly et al. (1987) dried 10 g of A. annua leaves in an oven at 40 ºC for 24 hours, then crushed and extracted with hexane (100 ml) by boiling for 15 minutes under reflux resulting in 0.153 % artemisinin. Marcel et al. (1997) compared supercritical fluid extraction (liquid carbon dioxide and ethanol in a ratio of (97:3)) resulting in 0.3% artemisinin and liquid solid extraction by ethyl acetate and found that extraction with ethyl acetate give more artemisinin (0.88 %). Parmeshwari and Ram, (1998) extracted artemisinin from 1.0 g of dried A. annua leaves in 15 ml hexane for 3 minutes at 40 ºC (0.009 % artemisinin). Kim et al. (2001) extracted 1.0 g roots of A. annua in 3 ml of ethyl acetate and hexane for 30 minutes at room temperature (0.003 % artemisinin). Guo et al. (2005) extracted 5 g of dried A. annua leaves in a Soxhlet extractor with 200 ml petroleum ether at 60 ºC for 6 hours (0.652 % artemisinin). Congyue et al. (2006) refluxed 0.5 g dried A. annua leaves in 50 ml of hexane at 75 ºC for 1 hour (0.0226 to 0.785 % artemisinin). Filip et al. (2006) has done extraction by immersing 1.0 gram of fresh plant material in 6 ml chloroform for one minute evaluating an artemisinin content of 0.068%. In the present study, extraction of artemether, arteether, artesunate and dihydroartemisinin from plant material was carried out using ethyl acetate and hexane in the ratio of 95:5 and acetonitrile 100% as it is most widely used laboratory techniques today (Van Quekelberghe et al., 2008; Yan Liu et al., 2009 ; Alexei et al., 2010). Ferreira et al., (1994) extracted artemisinin and its derivatives by refluxing 0.5 g of sieved dry leaves with 50 ml of hexane at 75 ◦C for 1 h. Artemether acid was extracted by many methods including extraction with hexane (Kim et al., 2001; Ferreira et al., 1994), chloroform (Mary et al., 1994) and with supercritical fluid composed of carbon dioxide and 3% methanol with temperature and pressure fixed at 50 degrees C and 15 MPa, respectively (Kohler et al., 1997). Arteether, artesunate and dihydro-artemisinin have been extracted mainly through hexane (Ferreira et al., 1994), Ethyl acetate (Ferreira and Janick, 2002), acetonitrile (Elsohly et al., 1987) and with chloroform (Woerdenbag et al., 1991; Mary et al., 1994). We used Ethyl Acetate because it increases solubility, and we use hexane for artemisinin because hexane and artemisinin both are non-polar, and non-polar dissolves in non-polar, when we add little amount of ethyl acetate into hexane, it increases the solubility of artemisinin Many analytical procedures to identify and quantify artemisinin, its biosynthetic precursors as well as its metabolites have been developed during the last 30 years. These include thin layer chromatography (TLC) (Klayman et al., 1984), High performance liquid chromatography with UV detection (HPLC-UV) (Zhao and Zeng, 1985), gas chromatography coupled to mass spectrometry (GC-MS) (Woerdenbag et al., 1991), capillary electrophoresis coupled to UV detector (CE-UV) (Christen and Veuthey, 2001) and enzyme-linked immunosorbent assay (ELIZA) (Jaziri et al., 1993). We successfully used HPLC equipped with auto-sampler, degasser and a photo-diode array detector method for rapid, accurate and cost effective detection and quantification of artemisinin and its derivatives. Now days, this is most commonly used method of many research groups for assay of artemisinin (Towler and weathers, 2007; Kim et al., 2001). Zhou et al. (1988) reported an HPLC method with polarographic detection of artemisinin and its derivatives. Several groups, (Sandrenan et al. (1997), Karbwang et al. (1997), Navaratnam et al. (1997), Bangchang et al., (1998) developed an analytical method for the determination of artemether, arteether, dihydro-artemisinin and artesunate based on HPLC and ECD in reductive. HPLC with evaporative light scattering detection (HPLC–ELSD) (Kohler et al., 1997; Christen and Veuthey, 2001), gas chromatography with mass spectrometric detection (GC– MS) (Woerdenbag et al., 1991), GC with flame ionization detection (GC–FID) (Ferreira et al., 1994; Sipahimalani et al., 1991) and enzyme-linked immunosorbent assay (ELISA) (Jaziri et al., 1993) are commonly used for analysis of artemether, arteether, artesunate and dihydro-artemisinin. Artemisinin, artemether, arteether, artesunate and dihydro-artemisinin analysis was carried out on dried plant parts (shoots, roots and hairy roots) of transformed and non-transformed Artemisia annua and Artemisia dubia for the present study. Similar type of dried plant parts of Artemisia species were used by a large number of other researchers for analysis of artemisinin and its derivatives (Liu et al., 1979; Acton et al., 1985; Klayman, 1985; ElSohly et al., 1987; Luo and Shen, 1987; Singh et al., 1988; Charles et al., 1990; Delabays et al., 1993; Woerdenbag et al., 1994; Marcel et al., 1997; Parmeshwari and Ram, 1998; Aryanti et al., 2001; Christen and Veuthey, 2001; Delabays et al., 2001; Guo et al., 2005; Mingfu et al., 2005; Arab et al., 2006; Congyue et al., 2006; Towler and Weathers, 2007). A prominent feature of our research work is to screen out both transformed and non-transformed Artemisia species for the presence of artemisinin and its derivatives by using same conditions of extraction (using ethyl acetate and hexane in the ratio of 95:5 and 100% acetonitrile), same conditions and equipment for analysis (HPLC equipped with auto-sampler, degasser and a photo-diode array detector) and same type of plant parts (dried plant parts). Complete data were analyzed by using same type of software (ANOVA). Therefore, all comparative results can be considered highly reliable and reproducible. Artemisinin analysis was carried out on dried plant material (shoots, roots and hairy roots) of transformed and non-transformed plants of A. annua and A. dubia. The tissue assayed was similar to that from Artemisia species in previous studies (Delabays et al.,2001; Guo et al., 2005; Mingfu et al., 2005; Arab et al., 2006; Congyue et al., 2006; Towler and Weathers, 2007). Our control plants contained similar amounts of artemisinin when compared to these studies on the vegetative stage. We found 0.4% DW artemisinin in control plants of A.annua and 0.01 % DW artemisinin in A.dubia, while Wallaart (1999) et al., found 0.1-0.5% DW artemisinin in control plants of A.annua, Weathers et al., (2007) found 0.2 -0.5% artemisinin in vegetative leaves and Mannan et al., (2010) found 0.44% DW artemisinin in A.annua leaves and 0.05% DW in leaves of A.dubia. Transformed shoots, roots of transformed shoots and hairy roots of A. annua and A. dubia all contained significantly more artemisinin when compared to the controls. The highest concentration of artemisinin was observed in transgenic leaves of both Artemisia species (i.e. ~ 10 and ~ 30 fold increase in leaves of A. annua and A. dubia respectively compared to controls). Other researchers have previously reported that transformation of different Artemisia species results in an increase in production of artemisinin within the plant. Vergauwe et al., (1998) reported that Agrobacterium tumefacienes transformation with EHA101 (pTJK136) strain and C58C1RifR (pGV2260) (pGSJ780BAT) strain of A. annua resulted in an increase in artemisinin content (0.42%) and its metabolites. Another report by He et al., (1983) reported 0.36% increase in artemisinin amount in leaves transformed with ADS gene of artemisinin biosynthetic pathway of A. annua as compared to non-transformed plants. Similarly, Abdin et al. (2003) reported the presence of artemisinin, ranging from 0.01 % to 0.80 %, in leaves and flowers of transformed A. annua and by overexpression of genes coding for enzymes associated with the rate limiting steps of artemisinin biosynthesis pathway. Han et al. (2006) reported that artemisinin production increased by up regulation of FDS through genetic transformation of Artemisia annua with different genes. Chen et al. (1999, 2000) described that Overexpression of Gossypium arboretum FPS in hairy roots and transgenic plants resulted in three to four fold and two to three folds increased in artemisinin content respectively. The results achieved in this study are significantly higher with increase of 12 fold or greater being detected depending on the tissue examined. A meaningful increase in artemisinin content was observed in hairy roots and roots of transformed shoots of A.annua and A.dubia as compared to control. Similarly roots of transformed plants in both Artemisia species also showed significant increase in artemisinin content compared to control roots. An earlier report shows that A. rhizogenes mediated transformation increased the level of secondary metabolites in plants (Kovalenko et al., 2004). Artemisinin production (0.42%) has been reported in hairy roots culture of A. annua (Weathers et al., 1994). Mannan et al., (2008) reported the production of artemisinin by hairy roots in Artemisia dubia (0.603%) and Artemisia indica (0.0026%) with Agrobacterium rhizogenes strain LBA9402. Weathers et al., (2005) reported a significant increase in artemisinin content in hairy roots culture of A.annua. The concentration of artemisinin in roots of untransformed A.annua and A.dubia is almost negligible (Figure 3.3). These findings agree with those of Kim et al. (2001) and Ferreira and Janick, (1996), who found very low artemisinin concentrations (0.003 %) in root of A. annua. Other studies did not detect artemisinin in roots of A. annua perhaps due to the sensitivity of the analysis carried out (He et al., 1983; Kudakasseril, 1987; Fulzele et al., 1991; Woerdenbag et al., 1991; Kim and Kim, 1992; Gulati et al., 1996). We also found low (A.annua) or zero (A.dubia) artemisinin concentration in untransformed roots. Transformed plants, hairy roots and roots of transformed shoots of A.dubia and A.annua were all found to contain significantly higher amounts of artemisinin. Our results clearly show that the roots of transformed plants and hairy roots both contain around 1.5% artemisinin. This is only slightly less than that found in leaves of A.annua and significantly higher than that found in leaf material from A.dubia. It has been reported that exogenous artemisinin is capable of inhibiting root growth at concentrations greater than 10 µg ml-1 (Arsenault et al., 2010). Our results showed that hairy roots were capable of producing more than 15ug g-1 DW. This would suggest that there are certain cells within the roots which are acting as storage compartments for the artemisinin produced. Mannan et al. (2010), has shown that roots of Artemisia plants are capable of producing artemisinin. Interestingly the roots do not contain trichomes; hair cells are however present and it is possible that these cells are also involved in the production and storage of artemisinin. The number of hair cells also appeared to be present in greater abundance on the leaves of transformed plants (Figure 4.2B). Artemether, arteether, artesunate and dihydro-artemisinin analysis was carried out on dried plant material (shoots, roots and hairy roots) of transformed and nontransformed plants of A. annua and A. dubia. The tissue assayed was similar to that from Artemisia species in previous studies (Delabays et al.,2001; Guo et al., 2005; Mingfu et al., 2005; Arab et al., 2006; Congyue et al., 2006; Towler and Weathers, 2007). Our control plants contained 0.105% DW of artemether, 0.15% DW of arteether, 0.104 % DW of artesunate and 0.2% DW of dihydro-artemisinin in control plants of A.annua and 0.03 % DW of artemether, 0.015 % DW of arteether, 0.013% DW of artesunate and 0.02% DW of dihydro-artemisinin in A.dubia. Transformed shoots, roots of transformed shoots and hairy roots of A. annua and A. dubia all contained significantly more amount of artemether, arteether, artesunate and dihydro-artemisinin when compared to the controls. The highest concentration of these artemisinin derivatives were observed in transgenic leaves of both Artemisia species (i.e. ~ 12, ~ 20, ~ 14, ~ 11 fold increase in artemether, arteether, artesunate and dihydro-artemisinin respectively in leaves of A. annua, and ~ 20, ~ 90, ~ 60, ~ 50 fold increase in artemether, arteether, artesunate and dihydro-artemisinin respectively ~ 30 fold increase in leaves of A. dubia compared to controls. To the best of our knowledge, there is no previous report about effect of transformation on enhanced production of derivatives of artemisinin. This is first report explaining the effect of transformation on enhanced production of artemether, arteether, artesunate and dihydro-artemisinin along with increased production of artemisinin in different tissues of Artemisia annua and Artemisia dubia. All the peaks in HPLC were confirmed by using spiked sample in each case for analysis of artemisinin and its derivatives. In some cases like artemisinin peak of standard was not matched with the sample because of shifting retention time due to Annua's matrix and that is why we used the spiked sample which showed increase in the artemisinin peak. This was one of the reasons for spiking. In general, statistical analysis showed that artemisinin content was enhanced much more as compared to its derivatives under the effect of transformation with rol genes and among two Artemisia species, A.annua is the best species for the production of artemisinin and its derivatives. Considering the tissue types, shoots and hairy roots are more appropriate for production of artemisinin and its derivatives. 7.4: Analysis of metabolic pathway and trichome development In the present study, expression of all the genes involved in artemisinin synthesis pathway were found to be significantly increased in all transgenic lines of A.dubia and A.annua (Fig.4.1A, 4.1B, 4.1C, 4.1D and 4.1E), which is in clear agreement with the data obtained for artemisinin content. Expression of ADS and CYP71AV1, of which the corresponding proteins catalyze the formation of amorpha-4, 11-diene and its ultimate conversion to dihydroartemisinic acid, was increased significantly in transformed shoots of A.annua and A.dubia. Similarly significant differences in expression of these genes were observed in hairy roots and roots of transformed plants of both A.annua and A.dubia compared to untransformed roots. The expression of CYP71AVI corresponds to that found by Olofsson et al., (2011) however the result for the ADS gene differs as they were unable to show increase in ADS gene expression in hairy roots. The ALDH1 gene which catalyses the oxidation of artemisinic and dihydroartemisinic aldehydes was also highly expressed in transformed plants of both Artemisia species as compared to untransformed plants, i.e. ~ 7 and ~ 5 times more in transformed leaves of A.annua and A.dubia, respectively, ~ 5times in roots of transformed shoots of A.annua and in A.dubia this increased from negligible amounts to significant levels, with similar increases in hairy roots compared to untransformed roots (Fig.4.1A, 4.1B, 4.1C and 4.1D). Although all the genes studied showed increased expression in transformed plants and hairy roots, however induction of CYP71AV1 and ALDH1 was much more compared to the ADS gene. TFAR1, known to stimulate trichome development and to catalyze sesquiterpenoid biosynthesis (Lies et al., 2011) showed significantly increased expression in all transgenic lines of A.annua and A.dubia (Fig. 4.2A and 4.2B). We also measured the density of trichomes in transformed and non-transformed plants and found that transformed leaves of A.annua and A.dubia have a higher density of trichomes compared to leaves of non-transformed plants. Despite close examination no trichomes were found on roots and hairy roots cultures of either Artemisia species but they were shown to contain significant amount of artemisinin as shown in Fig. 3.1. Olofsson et al., (2011), found that leaves and flower buds of A.annua produce more trichomes than other parts of the plant and also found that roots do not have trichomes. Although the pathway for the synthesis of artemisinin is well known (Schramek et al., 2010), it has been widely assumed that the trichomes are the only source of artemisinin production in the plant (Ferreira and Janick, 1995, 1996b; Tellez et al., 1999; Olsson et al., 2009) despite several papers describing its presence within roots (Mannan et al, 2010; Jaziri et al., 1995; Nair et al., 1986; Jha et al., 1988). Lommen et al (2006) reported that the artemisinin content of the plant is associated with a rise in trichome density but it continued to rise even after a collapse in the trichome populations. Our results found the trichome density in A. annua is increased by 10% following transformation however the artemisinin content increased by approximately 10 times. In A.dubia trichome density was increased by 9% following transformation and artemisinin content was increased by approximately 30 times (Fig.4.2A, 4.2B and Fig. 3.1). These results suggest a possible dissociation between trichome density and artemisinin content. It is unclear if the genes involved in artemisinin production are expressed elsewhere in the leaves (Duke et al, 1994). These results also suggest that other genes may involved in trichome development because TFAR1 gene also showed expression in roots and hairy roots but these did not produce any trichomes, or TFAR1 may also require some other genes for trichome development which may be not present in roots and hairy roots. Despite the work that has been carried out the maximum artemisinin content that has been achieved through natural selection is approximately 2% (Graham et al, 2010). It has been previously suggested that the plant was unable to support concentrations higher than this and that the trichomes capacity for storing artemisinin is limited (Liu et al., 2010). The current study shows that the plants are capable of producing and storing much higher amounts of artemisinin than previously thought both in leaves and in roots. This study shows significant amount of artemisinin production in transformed roots of A.annua and A.dubia and a small amount of artemisinin in untransformed roots of A.annua. The detection of significantly raised levels of expression of the genes involved in artemisinin in transformed roots associated with the detection of significant amounts of artemisinin in these tissues suggest that synthesis is occurring elsewhere in the plant as suggested by Duke et al., (1994). It should now be possible to determine how roots produce artemisinin hence elucidating how production could be improved. Baldi and Dixit (2008) and Wang et al. (2010) have reported that artemisinin production in plants and culture cells is increased after treatment with methyl Jasmonate; another report by Lies et al. (2011) examined the effect of cytokinin, gibberelin and jasmonate on artemisinin production. Similarly Patrick et al. (2010) described the effect of sugars on the production of artemisinin in A.annua. We transformed A.annua and A.dubia plants with rol ABC genes of A.tumefaciens and A.rhizogenes which are responsible for enhancing production of secondary metabolites in plants. The exact mechanism for the action of the rol genes is still far from clear (Bulgakov, 2008). It has been suggested that they act through the stimulation of the plants deference response which includes induction of many of the hormonal pathways. Many secondary metabolites are thought to be have evolved to protect plants as they expanded to fill different evolutionary niches. It has been suggested that artemisinin is an example of such a metabolite as it acts to prevent predation of the plant by herbivores and it is stimulated by pruning prior to harvest (Duke et al., 1987). This may explain why the transformation with Agrobacterium rol genes has such a significant effect on the amount of artemisinin produced by the transformed plants. The detection of significantly raised levels of expression of the genes involved in artemisinin biosynthesis in transformed roots can be correlated with the production of significant amounts of artemisinin in root tissues. This suggests that synthesis is occurring in tissues other than the trichomes which contradict previous theories and alter our perception of how and where artemisinin is produced within the plants. These findings will help to develop better strategies to increase the production of this valuable therapeutic drug which will in turn allow greater use of it in the chemotherapy of malaria and other diseases. 7.5: Analysis of transformed and untransformed Artemisia annua and Artemisia dubia on breast cancer cell lines Breast cancer is the most frequent type of cancer all over the world and major source of mortality among women. Schmidt et al., (2007) reported that death rate from breast cancer is about 38 deaths per 100,000 women, although a great improvement in breast cancer treatment has been made and decrease in the death rate is also observed in advanced countries. MCF-7 is a breast cancer cell line isolated from 69 years old Caucasian women and established by Soule et al. (1973). MCF-7 is estrogens receptor positive breast cancer cell lines. To overcome this cancer a constant research addressing more selective natural inhibitors with low side effects is continued and each discovery adds to the target. It has been documented that artemisinin inhibits cell proliferation in breast cancer and other types of cancers (Tan et al., 2011). In the current project, the effect of crude extracts of transformed and untransformed plants of Artemisia annua and Artemisia dubia on MCF-7 breast cancer cell lines was studied by using method of Vanicha and Kanyawim (2006) Sulforhodamine B (SRB) antiproliferative assay. In the present study maximum proliferation inhibition at 50µg was observed in hexane extracts of transformed plants of A.annua (98%) and A.dubia (93%) respectively with lower IC50 values compared to control plants (Fig.5.1A) with higher IC50 values. Lai and Sing (2006) also reported that artemisinin is very effective against breast cancer; they treated rats with 0.02 % artemisinin. After 40 weeks the rats were monitored for breast cancer. Oral artemisinin significantly delayed (p <0.02) and in some animals prevented breast cancer development, i.e. 57 % of artemisininfed versus 96 % of the controls. Breast tumors in artemisinin rats were fewer (p <0.02) and smaller in size (p < 0.05) when compared with controls. Efferth (2006) also explained the effect of artemisinin and its derivatives on several cancer cells and reported that artemisinin is very effective drug against many types of cancers especially breast cancer, colon cancer and lung carcinomas without any side effects. Similarly significant proliferation inhibition was observed in hairy roots of both A.annua (91-93%) and A.dubia (85%) compared to untransformed roots (Fig. 5.1A). The aqueous extracts of transformed and untransformed plants of A.dubia and A.annua were also tested on MCF-7 breast cancer cells and it was found that transformed plants of both Artemisia species (89% in A.annua and 84% in A.dubia) and hairy roots (83% of A.annua and 79% of A.dubia) showed more proliferation inhibition compared to control plants and roots respectively (Fig. 5.1B). Many researchers (Efferth et al., 2001; Sing and Lai 2001) reported the antitumor and cytotoxic potential of crude methanolic extract of A. annua and its n-hexane and aqueous fractions, using potato disc antitumor assay and SRB cytotoxicity assay using H157 and HT144 human tumor cell lines. They reported that the aqueous fractions had most significant activity among all the tested samples. The MCF-7 human cancer cell line was tested and hexane fraction was found show higher proliferation inhibition of MCF-7 at 50µM concentration suggesting that A. annua extract had selective inhibitory potential against different cell lines. Furthermore, the compounds with selective inhibitory potential can be isolated from n-hexane and aqueous fractions. Overall results and IC50 values showed that transgenic and non-transgenic plants of Artemisia annua and Artemisia dubia have significant activity in antiproliferative assay compared to aqueous extracts of Artemisia annua and Artemisia dubia plants (Fig. 5.1 ABC). As artemisinin is better dissolved in hexane than water due to its physico-chemical properties (low thermal and chemical stability of the endoperoxide function, low polarity and, hence, poor solubility in water and good solubility in organic solvents) (Lapkin and Plucinski, 2006; Wright et al., 2010). It confirmed the effect of artemisinin on cancer cells compared to other secondary metabolites present in transformed and untransformed Artemisia annua and Artemisia dubia because artemisinin content and anticancer activity showed significant correlation with each other (Fig 5.2) in case of hexane extraction compared to aqueous extraction. Higher proliferation inhibition was observed in transgenic plants and hairy roots of Artemisia annua as compared to transgenic plants and hairy roots of Artemisia dubia because A.annua plants produced more artemisinin compared to A.dubia plants (Fig. 5.1C). Various other researchers have also reported the potential anticancer properties of artemisinin and its analogs (Beekman et al., 1997, 1998; Chen et al., 2003, 2004; Efferth et al., 2001; 2002; Efferth and Oesch, 2004; Jeyadevan et al., 2004; Lee et al., 2000; Li et al., 2001; Mukanganyama et al., 2002; Posner et al., 1999, 2003, 2004; Reungpatthanapong and Mankhetkorn, 2002; Sadava et al., 2002; Sun et al., 1992; Woerdenbag et al., 1993, Wu et al., 2001) but at the same time we cannot skip the affectivity of other compounds present in crude extracts of both Artemisia species. Overall these results showed that the Artemisia plants from Pakistan might be a promising source of pharmacologically active natural products and among these Artemisia annua might be the most promising plant for further studies as the extracts/fractions from this plant showed highest and significant activity. This plant has the widest use in malaria treatment from many years and now for treatment of cancer among all the tested species of Artemisia. This suggests that the bioactive compounds isolated from this plant will be safe for human consumption as medicine and hence splashes our interest for further studies. Therefore, it can be concluded that Artemisia species had promising potential for further isolation and purification of cancer chemopreventive and anticancer compounds. Furthermore, anticancerous activity could be increased with the enhanced production of other secondary metabolites including artemisinin in rol genes transformed plants of Artemisia annua and Artemisia dubia. 7.6: Sequence analysis of Artemisia annua and Artemisia dubia and comparison in transformed and untransformed plants Different species evolved as a result of speciation event during the course of evolution (John and Jeroen, 2004; Ohta, 1989). We come across increasingly complex genomes as we move up the evolutionary tree. There can be several reasons that have caused the complexity of genome. Genes can be acquired from other species or they can be duplicated from the existing genes. Duplication of single gene or group families (Ohta, 2000), entire genome or a single chromosome or part of it can result in large genome sequences. Duplication of entire genome is the most rapid means of increasing gene expression (Regis et al., 2009) and number. Sequences that emerged as a result of recombination within the genome or were carried across the species by any means lead to divergence of genomes (Posada et al., 2002). As for example, viruses can integrate themselves in the host genome. They can incorporate their own genes and also those that they carry across from different species. The fact that viruses can cross the species barrier makes it even attractive proposition (Linda and Van, 2009). Replication slippage and DNA amplification (David and Rafael, 1997) mostly create gene duplication. These evolutionary tools lead to sequence homology amongst different organisms. Their genomes overlap to varying degree depending upon the time of divergence. Proteins in different species show shared features. Consensus sequences give a rational proof of structural and fictional similarity of proteins. In this study, the transcriptome of Artemisia annua and Artemisia dubia were sequenced for the first time. Gene contigs were generated by alignment of the sequences derived from a compilation of all the sequences obtained from a species. The different individual samples from both A.annua and A.dubia were then compared against this reference gene list to evaluate if any genes in the genome of the transformed plants are up or down regulated compared to untransformed plants. The Bowtie and CuffLinks programs were used to look for differential expression and CuffDiff used to derive values. A comparison was also made between A.annua and A.dubia (Table 6.1) plants. Contigs/genes for which significant differences were obtained were compared between biological replicates to determine their overall significance. Those which showed significant differences were submitted to the BLAST program using both the nr/nt and EST database available at NCBI. The homology of their genome sequence was found and significant results were obtained. From the results obtained the putative gene functions could be derived. Furthermore, genome sequences of individual transgenic lines, which were different from each other in copy number, were also compared. Besides, genome sequence of untransformed and transformed A.annua was compared individually with genome sequence of both untransformed and transformed A.dubia (Table 6.2). Up regulation and down regulation of different samples was analyzed on the basis of significant value obtained by generating contigs from different samples (Table 6.1). Those samples were considered to be highly up regulated that contained significant value less than 0.01. Transgenic lines of A.annua (A1, A2) and A.dubia (D1) were compared with control plants (Ac and Dc, respectively). Genes of both transgenic lines were found to be highly up regulated as compared to control plants. Although the different transgenic lines produced were inserted with only rol genes, the whole transcriptome was found upregulated. The reason behind this factor can be the influence that these rol genes might be able to exert on the expression system of other coding region of genome. Findings about the over expression of genes that had different copy number in transgenic lines (A1 and D1 has two copy number, A2 has single copy number) can be testified. Principally, those genes that have more copy number, shall exhibit more number of transcripts, conditioning that both copies are functional to produce transcripts. Another factor can be the modulation of promoter region in such a way that may have played role in the enhanced expression of gene. Inserted rol genes construct can integrate in gene sequence to cause Insertional mutagenesis (Yi Hong, 2008; Regis et al., 2009). It can also interfere with the promoter region and affect the expression (Renata et al., 2010). The sequence of rol ABC gene may have provided sites for epigenetic modification that has altered the DNA sequence to assist the transcription process (Austin and Dixon, 1992; Grønbaek et al., 2007; Riggs et al., 1996; Portela and Esteller, 2010). The sequence of inserted rol gene can also play role as an activator or enhancer. When two transgenic lines of A.annua (A1 and A2) were compared with each other, transgenic line carrying two copy number (A1) was found up regulated compared to transgenic line having one copy number (A2). It shows that both the copies of A1 transgenic line were functional and giving enhanced expression relative to A2 with single copy of gene confirming the findings of Jun et al., (2011) who explained that gene expression is directly related to the copy number of genes. Transformed and untransformed plants of A.annua when compared with transformed plants of A.dubia were shown upregulated. Up regulation was also observed when transformed and untransformed plants of A.annua were compared with control plants of A.dubia. These results show that genome of A.annua is more prone to epigenetic modifications (Portela and Esteller, 2010). In A.annua Insulator sequences seem more efficient in molecular assembly on them (Maria et al., 2012) and are a route of influence to give it an edge for up regulation over transgenic plants of A.dubia. It is interesting to note that the pattern of artemisinin production happens to match with the up regulation of genes of transgenic lines shown in table 6.1. Amount of artemisinin was found higher in transgenic lines of A.annua as compared to control. The same pattern follows when gene up regulation is analyzed between transgenic and control plants (Table 6.1). Similarly, transgenic lines of A.annua and A.dubia were examined for artemisinin production, transformed A.annua showed more amount of artemisinin than transgenic line of A.dubia (Graph 3.1). Gene expression pattern of A.annua and A.dubia transgenic lines goes in harmony with the artemisinin production in both lines. Same can be observed by analysis of control and transgenic lines of A.dubia, where transformed samples have more artemisinin content, a pattern that is similar to gene up regulation in these two lines. Hence gene up regulation and artemisinin production can be correlated. It can be said that rol genes affect the genes of artemisinin biosynthesis in a similar fashion as they do to up regulate other genes. This speculation makes sense because it is represented by our results of RT-PCR (Fig 4.1 A, B, C, D, and E) that rol genes affect the genes involved in artemisinin biosynthesis pathway. BLAST of Various contigs which were obtained by aligning different samples of A.annua and A.dubia were performed and sequence homology presented interesting results. These contigs showed homology with different genome sequences ranging from 60% to 90% but only those Contigs were analyzed that showed 90% homology or more than it (Table 6.2). Contigs found as a result of aligning A1 and A2 produced more than 90% homology with different genes of Lactuca sativa, Solanum lycopersicum, Lycopersicum esculentum, Vitis vinifera and Malus x domestica. Amongst those genes WRKY 17 which is a key regulatory transcription factor in the artemisinin biosynthesis pathway (Dongming et al., 2009) was also found with the highest similarity to Malus x domestica (Apple). Hence it can be suggested that in Malus x domestica WKRY 17 influences pathway; it is interesting to speculate what pathway this might be. Like A1 and A2, contigs from other samples of A.annua and A. dubia (A1 and AC, A2 and AC, A1 and D1, A2 and D1, AC and D1, A1 and DC, A2 and DC, D1 and DC and AC and DC) also showed sequence homology with different genes of different genomes present in nr/nt and EST databases of NCBI. These contigs showed 97%-100% sequence homology mostly with genomes of Homo sapiens, Mus musculus, Arabidopsis thaliana, Drosophila, Vitis vinifera, Oryza sativa, Lycopersicum esculentum, Coffea arabica, Solanum lycopersicum, Mediacago truncatula, Helianthus species, Honey bee, Zebra fish, Zea mays, pig and some EST sequences of A.annua. This sequence homology of contigs may suggest that these genes bear functional similarity (Brown, 2002). It is also possible that this overlapping sequence may code a part of transcript that is common to them (Eugene and Michael, 2003). It can also mean that the overlapping sequences are evolutionarily closely related that diverged over the course of time (Koonin, 2005). The presence of almost 100% sequence homology of contigs from these two Artemisia species with different animals e.g. Homo sapiens, Mus musculus, Honey bee, Zebra fish and Pig may represent the presence of any gene that codes for product which is common in all species. Presence of similar sequences in contemporary plants and animals support the idea of common progenitor (Maria et al., 1990). Conclusion Conclusively, this project represents the comparison of the artemisinin contents observed in the transformed and untransformed shoots, roots and hairy roots of Artemisia annua and Artemisia dubia transformed with rol genes through A.tumefacienes and A.rhizogenes mediated transformation. Artemisinin was significantly increased in the transformed shoots, hairy roots and roots of transformed shoots compared to control in both A.annua and A.dubia plants. Artemether, arteether, dihydroartemisinin and artesunate contents were also significantly increased in transgenic shoots, roots and hairy roots to control. Amount found in roots of transformed shoots and transformed hairy roots was significantly greater compared to the control roots in which negligible amounts was detected. Metabolic pathway of artemisinin production in Artemisia annua and Artemisia dubia and the effect of rol genes through which these genes enhance the production of artemisinin were analyzed using different genes involved in artemisinin production and TFAR1 involved in trichome development and sesquiterpenoid biosynthesis was also studied. Pronounced changes in the expression of the artemisinin biosynthetic pathway genes were observed in all transgenic lines of A. annua and A. dubia transformed with rol ABC genes. Expression levels of ADS, CYP71AV1, ALDH1 and TFAR1 were significantly increased in transformed plants and hairy roots of both Artemisia species compared to untransformed plants. Hexane and aqueous extracts from transformed and untransformed plants of Artemisia annua and Artemisia dubia were analyzed for anticancerous activity against MCF-7 breast cancer cell lines. Crude extracts of rol genes transgenic plants revealed higher anticancerous activities against MCF-7 breast cancer cell compared to control plants. Hexane extracts of rol ABC genes transgenic plants revealed higher anticancerous activity against MCF-7 breast cancer cell lines compared to aqueous extracts of plants. Transcriptome of Artemisia annua and Artemisia dubia were sequenced for the first time. Genes of transgenic lines of Artemisia annua and Artemisia dubia were found to be highly up regulated as compared to control plants. Moreover, transgenic lines carrying two copy numbers were found up regulated compared to transgenic lines having one copy number. BLAST of different contigs also showed many enzymes that have established role in A.annua and A. dubia and were common to various other species. These results can have many implications and can be extrapolated to understand different unknown aspects for a newly sequenced genome. Future strategies  Transformation with individual rol A, B and C genes could be carried out to check their effect on artemisinin production and their comparison with combined rol ABC genes could be studied.  Further research may be carried out to study the gene expression in T1 and T2 generations by RT-PCR and Northern blot analysis.  Maximum number of Artemisia species could be genetically transformed with rol ABC genes for increased, constant and permanent source of this valuable drug.  Artemisinin content could be further increase by transformation with other genes involved in artemisinin biosynthesis pathway.  Other In-vivo and In-vitro assays may be carried out to unlock other properties of artemisinin and different Artemisia species.  Sequencing of whole genome of Artemisia may be helpful in predicting the molecular functions of genes of different Artemisia species.  Stimulation of the synthesis of artemisinin within these plants has also allowed us to alter our perception of how and where artemisinin is produced within the plants. These findings will help to develop better strategies to increase the production of this valuable therapeutic drug which will in turn allow greater use of it in the chemotherapy of malaria and other diseases. REFRENCES References Abdalla SS, Abu-Zarga MH (1987). Effect of cirsimaritin, flavones isolated from Artemisia Judaica, an isolate guinea-pigilum. Planta Med. 53: 322-324. Abdin MZ, Israr M, Rehman RU, Jain,SK (2003). Artemisinin, a novel antimalarial drug: biochemical and molecular approaches for enhanced production. Planta Med. 69: 289-299. Acton N, Klayman DL, Rollmann IJ (1985). Reductive electrochemical HPLC assay for artemisinin (qinghaosu). Planta Med. 5: 445-446. Acton N, Klayman DL, Rollmann IJ, Novotny JF (1986). Isolation of artemisinin (qinghaosu) and its separation from artemisitene using its multilayer coil separator-extractor and isolation of arteannuin B. J. chromatogr. 355: 448-450. Adam KP, Thiel R, Zapp J, Becker H (1998). Involvement of the mevalonic acid pathway and the glyceraldehydes-pyruvate pathway in terpenoid biosynthesis of the liverworts Ricciocarpos natans and Conocephalum conicum. Arch. Biochem. Biophys. 354: 181-187. Ahn S, Anderson JA, Sorrells ME, Tanksley SD (1993). Homoeologous relationships of rice, wheat and maize chromosomes. Mol. Gen. Genet. 241, 483–490. Alexei AL, Martina P, Lasse G, Smain C, Kai L, Marcel AL, Walter L (2010). Screening of new solvents for artemisinin extraction process using ab initio methodology. Green Chemistry. 2 (12). DOI: 10.1039/b922001a. Altabella T, Angel E, Biondi S, Palazon J, Bagni N (1995). Effect of the rol genes from Agrobacterium rhizogenes on palyomine metabolism in tobacco roots. Phys. Plant. 95: 479-485. Altamura, MM, Capitani F, Gazza L, Capone I, Castantino P (1998). The plant oncogene rol β stimulates the flower and root meristemoid in tobacco cell layers. Wew phytol. 126: 283-298. Altvorst AC, Bino RJ, Van Dijk AJ, Lamers AMJ, Van Den Mark F, Dons JJM (1992). Effects of the introduction of Agrobacteirum rhizogenes rol genes on Artemisia plant and flower development. Plant Sci. 83: 77-85. Anderson GK (1977). Indian medicinal plant (2nd Ed.). Blater, E., Caius, J. F. and Makaskar K. S. (Eds.). International book distributors (India). Pp. 1391-1402. Andre F (2000). Hepatitis B epidemiology in Asia, the Middle East and Africa. Vaccine 18(1): 20-22. Anon (1979). Qinghaosu antimalarial coordinating research group. Antimalarial study on ginghaosu. Chin. Med. J. 92: 811-816. Anon (1981). Fourth meeting of the scientific working group on the chemotherapy of Malaria, Beijing, People‟s Republic of China. WHO Reports. TDR / CHEMAL - SWG (4) / QHS/813. p. 5. Anon (1982). China Cooperative Research Group on Qinghaosu and its Derivatives as antimalarials. J. Trad. Chin. Med. 2: 3-8. Anon (1992). Rediscovering wormwood, Qinghaosu for malaria. Lancet 339: 649651. Arab HA, Rahbari S, Rassouli A, Moslemi MH (2006). Khosravirad FDA: Determination of artemisinin in Artemisia sieberi and anticoccidial effects of the plant extract in broiler chickens. Trop Anim Health Prod. 38:497-503. Arino A, Arberas I, Renobales G, Arriaqa S, Dominquez JB (1999). Essential oils of Artemisia absinthium L. from the Spanish Pyrenees. J. Ess. Oil Resid. 11(2): 182-184. Arsenault PR, Vail DR, Wobbe KK, Weathers PJ (2010). Effect of sugars on artemisinin production in Artemisia annua L. transcription and metabolite measurements. Molecules. 15: 2302–2318. Aryanti, Bintang M, Ermayanti TM, Mariska I (2001). Production of antileukemic agent in untransformed and transformed roots cultures of Artemisia cina. Ann. Bogor. 8(1): 11-16. Austin S, Dixon R (1992). The prokaryotic enhancer binding protein NTRC has an ATPase activity which is phosphorylation and DNA dependent. EMBO J. 11(6): 2219–2228. PMCID: PMC556689. Avery MA, Muraleedharan KM, Desai PV, Bandyopadhyaya AK, Furtado MM, Tekwani BL (2003). Structure-activity relationships of the antimalarial agent artemisinin. 8. design, synthesis, and CoMFA studies toward the development of artemisinin-based drugs against leishmaniasis and malaria. J. Med. Chem. 46: 4244-4258 Bakkali AT, Jaziri M, Foriers A, Vander Heyden Y, Vanhaelen M, Homes J (1997). Lawsone accumulation in normal and transformed cultures of henna, Lawsonia inermis. Plant Cell Tissue Organ Cult. 51: 83-87. Baldi A, Dixit VK (2008) Yield enhancement strategies for artemisinin production by suspension culture of Artemisia annua. Bioresour Technol 99: 4609-4614 Balint G (2001). Artemisinin and its derivatives an important new class of antimalarial agents. Pharmacology & Therapeutics. 90: 261-265. Banerjee S, Rahman L, Uniyal GC, Ahuja PS (1998). Enhanced production of valepotriates by Agrobacterium rhizogenes induced hairy root cultures of Valeriana wallichii DC. Plant Sci. 131: 203-208. Bangchang K, Congpuong K, Hung LN, Molunto P, Karbwang J (1998). Simple high-performance liquid chromatographic method with electrochemical detection for the simultaneous determination of artesunate and dihydroartemisinin in biological fluids. J Chromatogr B Biomed Sci Appl. 24; 708(1-2):201-7. Banthorpe DV, Brown GD (1989). Two unexpected coumarin derivatives from tissue culture of compositae species. Phytochem. 28: 3003-3007. Baroncelli S, Buiatti M, Bennici A (1992). Genetics of growth and differentiation “in vitro” of Artemisia annua. Z.Pflanzenzuchtg. 70: 99-107. Basu BD, Kirtikar, KR (1988). Indian medicinal plant (2nd Ed.). Blater, E., Caius, J. F. and Makaskar K. S. (Eds.). International book distributors (India). Pp. 13911402. Beekman AC, Wierenga PK, Woerdenbag HJ, Uden WV, Pras ., Konings AWT, El-Feraly FS, Galal AM, Wikstrom HV (1998). Artemisinin-derived sesquiterpene lactones as potential antitumour compounds: cytotoxic action against bone marrow and tumour cells. Planta Medica 64, 615–619. Beekman AC, Woerdenbag HJ, Van Uden W, Pras N, Konings AWT, Wikstrom HV (1997). Stability of artemisinin in aqueous environments: impact on its cytotoxic action to Ehrlich ascites tumour cells. Journal of Pharmacy and Pharmacology 49, 1254–1258. Benjamin BD, Roja J, Heble, MR (1993). A. rhizogenes mediated transformation of Rauwalfia serpentina. Regeneration and alkaloid synthesis. Plant Cell tiss. Org. Cult. 35: 253-257 Bennett, MD, and Leitch IJ (1995). Nuclear DNA amounts in angiosperms. Ann. Bot. 76, 113–176. Bennetzen JL and Freeling M (1993). Grasses as a single genetic system: Genome composition, collinearity and compatibility. Trends Genet. 9, 259–261. Bertea CM, Freije JR, van der Woude H, Verstappen FWA, Perk L, Marquez V, De Kraker J-W, Posthumus MA, Jansen BJM, de Groot A (2005). Identification of intermediates and enzymes involved in the early steps of artemisinin biosynthesis in Artemisia annua. Planta Medica 71: 40–47. Bettini P, Michelotti S, Bindi D, Capuana M, buiatti M (2003). Pleotropic effect of the insertion of the A. rhizogenes rol D gene in tomato. Theor. Appl. Genet. 107(5): 831-836. Bidney DL, Shepard JF, Kaleikau E (1991). Regeneration of plants from different explants of Artemisia species. Plant. Cell. Tiss. 117: 89-92. Biswajit G, Swapna M, Sumita F, Gosh B, Jha S (2000). Genetic transformation of Artemisia annua by A. tumefacienes and artemisinin synthesis in transformed cultures. Plant Scio. Linerich. 122(2): 193-199. Biswajit G, Swapna M, Sumita J (1997). Genetic transformation of Artemisia annua by Agrobacterium tumefaciens and artemisinin synthesis in transformed cultures. Plant Science. 122 :193-199. Bonhomme V, LaurainMattar D, Fliniaux MA (2000a). Effects of the rolC gene on hairy root induction development and tropane alkaloid production by Atropa belladonna. J Nat Prod. 63:1249–52. Borrmann S, Szlezak N, Faucher JF, Matiegui PB, Neubauer R, Biner RK, Lell B, Kremsner PG (2001). Artesunate and praziquantel for the treatment of Schistosoma haematobium infections: a double-blind, randomixed, placebocontrolled study. J. Infect. Dis. 184: 1363-1366. Bremer K, Humphries C (1993). Generic monograph of the AsteraceaeAnthemideae. Bull. Nat. Hist. Museum, London 23: 1-177. Brown GD (1994). Secondary metabolism in tissue culture of Artemisia annua. J. Nat. Prod. 57(7):975-977. Brown GD, Sy L-K (2004). In vivo transformations of dihydroartemisinic acid in Artemisia annua plants. Tetrahedron 60: 1139–1159. Brown GD, Sy L-K (2007). In vivo transformations of artemisinic acid in Artemisia annua plants. Tetrahedron 63: 9548–9566. Brown TA (1993). Gene cloning and DNA analysis and introduction. Brown, T. A. (Ed.). Blackwell Science, pp.139-144. Brown TA (2001). Gene cloning and DNA analysis and introduction. Brown, T. A. (Ed.). Blackwell Science, pp.139-144. Brown TA (2002). Genomes, 2nd edition. ISBN-10: 0-471-25046-5. Bulgakov VP (2008). Functions of rol genes in plant secondary metabolism. Biotechnol Adv 26: 318-24. Bulgakov VP, Kusaykin M, Technenoded GK, Zuyagintseva TN, Zhuravlev YN (2002). Carbohydrase activities of the ral C gene ttransformed and nontransformed ginseng cultures. Fitoterapia. 73: 638-643. Bulgakov VP, Tchernoded GK, Mischenko NP, Khodakovskaya MV, Glazunov VP, Zvereva EV (2002a). Effects of salicylic acid, methyl jasmonate, etephone and cantharidin on anthraquinone production by Rubia cordifolia callus cultures transformed with rolBandrolC genes. J Biotechnol. 97:213–21. Bulgakov VP, Tchernoded GK, Mischenko NP, Shkryl YN, Glazunov VP, Fedoreyev SA (2003). Effects of Ca2+ channel blockers and protein kinase/phosphatase inhibitors on growth and anthraquinone production in Rubia cordifolia cultures transformed by the rol B and rolC genes. Planta. 217:349–55. Bulgakov VP, Tchernodeda GK, Mischenkob WP, Fedoreyevb SA, Zhuravleva XN (2003). The rol B and rol C genes activate synthesis of onthraquinones in Rubia cardifolia cells by mechanism independent of octaelecanoid signaling pathway. Plant Sci. 166: 1069-1075. Bundock P, den Dulk-Ras A, Beijersbergen A, Hooykass PJJ (2005). Transkingdom T-DNA transfer from agrobacterium to Saccharomyces cerevisial, Embo. J. 14(3): 3206. Caius JF (1986). Medicinal and Poisonous Plants of India. Scientific Publishers, Jodhpur, India p. 317-322. Caius JF, Mahaskar KS (1920). The correlation between the chemical composition of antihelmintics and their therapeutic values in connection with the hookworm inquiry in the Madras Presidency. II. Oleum chenopodii. III. Oleum absinthii. IV. Oleum tanaceti. Indian J. Med. Res. 7: 570-609. Camilleri C, Jauanin L (1999). The TR-DNA regin carrying the auxin synthesis genes of the A. rhizogenes agropine type plasmid pRi A4: nucleotide sequence analysis and introduction into tobacco plants. Mol. Plant Microbe Interact. 138: 103-109. Chan KL, Teo CK, Jinadasa S, Yuen KH (1995). Selection of high artemisinin yielding Artemisia annua. Planta Med. 61: 285-287. Chan MT, Chang HH, Ho SL, Tong WF, Yu SM (2004). Agrobacterium-mediated production of transgenic rice plants expressing a chimeric -amylase promoter/P-gluguronidase gene. Plant. Mol. Bial. 22: 491-506. Charles DJ, Cebert E, Simon JE (1991). Characterization of the essential oil of Artemisia annua L. J. Ess. Oil Res. 3: 33-39. Charles DJ, Simon JE (1990). Germplasm variation in artemisinin-content of Artemisia annua using an alternative method of artemisinin analysis from crude plant extracts. J. Nat. Prod. 53: 157-160. Chen D, Ye H, Li G (2000). Expression of a chimeric farnesyl diphosphate synthase gene in Artemisia annua L. transgenic plants via Agrobacterium tumefaciensmediated transformation.PlantSci. 155:179–185. Chen D, Ye H, Li G (2000a). Expression of a chimeric farnesyl diphosphate synthase gene in Artemisia annua L. transgenic plants via Agrobacterium tumafaciensmediated transformation. Plant Sci. 155: 179-185. Chen DH, Liu CJ, Ye HC, Li GF, Liu BY, Meng YL, Chen XY(1999). Rimediated transformation of Artemisia annua with a recombinant farnesyl diphosphate synthase gene for artemisinin production. Plant Cell Tissue Organ Cult. 57:157–162. Chen HH, Zhou HJ, Fang X (2003). Inhibition of human cancer cell line growth and human umbilical vein endothelial cell angiogenesis by artemisinin derivatives in vitro. Pharmacological Research 48, 231–236. Chen HH, Zhou HJ, Wu GD, Lou XE (2004). Inhibitory effects of artesunate on angiogenesis and on expressions of vascular endothelial growth factor and VEGF receptor KDR/flk-1. Pharmacology 71, 1–9. Chen PK, Leather GR (1990). Plant growth regulatory activities of artemisinin and its related compounds. J. Chem. Ecol. 16: 1867-1876. Chen PK, Polatnick M, Leather GR (1991). Comparative study on artemisinin, 2, 4-D, and glyphosate. J. Agric. Food. Chem. 39: 991-994. Chilton MD, Drummond MH, Gordon MP, Wester EW (2003). Stable incorporation of plasmid DNA into higher plant cells: the molecular basis of crown gall tumori genesis. Cell. 11: 268-271. Chilton MD, Topfer, DA, Petit A, David G, Cassa DF, Tenpe J (1982). A rhizogenes inserts T-DNA into the genomes of host plant root cells. Nature 295: 432-434. Christen P and Veuthey JL (2001). Veuthey, New trends in extraction, identification and quantification of artemisinin and its derivatives, Curr. Med. Chem. 8: 1827–1839. Clapham AR, Tutin TG, Warburg EF (1962). Flora of the British Isles. Cambridge University Press p. 855-862. Congyue AP, Jorge FS, Andrew JW (2006). Direct analysis of artemisinin from Artemisia annua L. using high-performance liquid chromatography with evaporative light scattering detector and gas chromatography with flame ionization detector. J. Chromat. A. 1133: 254-258. Covello PS (2008). Making artemisinin. Phytochemistry. 69: 2881–2885. Covello PS, Teoh KH, Polichuk DR, Reed DW, Nowak G (2007). Functional genomics and the biosynthesis of artemisinin. Phytochemistry 68: 1864–1871. Cronquist A (1955). Phylogeny and taxonomy of the compositae. Am. Midl. Nat. 53: 478-511. Cronquist A (1988). Asterales. In: Cronquist A, Holmgren A, Holmgren N, Reveal J, Holmgren P (eds), Intermountain Flora: vascular plants of the intermountain west, Vol. 5. New York, Hafner p. 543-560. Croteau R, Kutchan T, Lewis N (2000). Natural products (secondary metabolites). In: Buchanan B, Gruissem W, Jones R (eds), Biochemistry and molecular biology of plants. Am. Soc. Plant Physiol. Rockville, MD. 1250-1318. Cullen J (1975). Flora of Turkey and East Aegeam Islands. Davis, P. H. (Ed.). Edinburgh University Press, Edinburgh. 5: 311-325. Dasture JF (1952). Medicinal plants of India and Pakistan. Taraporevala Sons & Co. Ltd. Bombay. Pp.39-40. David R and Rafael P (1997). Gene amplification and genomic plasticity in prokaryotes. Annual review 10.1146/annurev.genet.31.1.91. of genetics. Vol. 31: 91-111. DOI: Davies MJ, Atkinson CJ, Burns C, Woolley JG, Hipps NA, Arroo RRJ, Dungey N, Robinson T, Brown P, Flockart I (2009). Enhancement of artemisinin concentration and yield in response to optimization of nitrogen and potassium supply to Artemisia annua. Annals of Botany. 104: 315–323. Davioud JV, Shank KM, Montagu MJC, Zambryski R (1988). Efficient and genotype independent Agrobacterium mediated transformation of Artemisia annua. J Plant Physiol. 160(10): 1253-1257. Davis TME, Karunajeewa HA, Illett KF (2005). Artemisinin-based combination therapies for uncomplicated malaria. Med. J. Aust. 182 (4), 181-185. Dehio C, Grossmann K, shell J, Schmulling T (1998). Phenotype and hormonal status of transgenic tobacco plants over expressing the rol A. gene of A. rhizogenes TL-DNA. Plant Mol. Biol. 23: 1199-1210. Delabays N, Benakis A, Collet G (1993). Selection and breeding for high artemisinin (Qinghaosu) yielding strains of Artemisia annua. Acta. Hort. 330: 203-207. Delabays N, Simonnet X, Gaudin M (2001). The genetics of artemisinin content in Artemisia annua L. and the breeding of high yielding cultivars. Current Medicinal Chemistry. 8: 1795–1801. Dell’Eva R, Pfeffer U, Vene R (2004). Inhibition of angiogenesis in vivo and growth of Kaposi‘s sarcoma xenograft tumors by the anti-malarial artesunate. Biochem Pharmacol. 68: 2359–66. Dellicour S, Hall S, Chandramohan D, Greenwood B (2007). The safety of artemisinins during pregnancy: a pressing question. Malar. J. 6: 15. Dhingra V, Rajoli C, Narasu ML (2000). Partial purification of proteins involved in the bioconversion of arteannuin B to artemisinin. Bioresour. Tech. 73: 279282. Disbrow GL, Baege AC, Kierpiec KA (2005). Dihydroartemisinin is Cytotoxic to papillomavirus-expressing epithelial cells in vitro and in vivo. Cancer Res. 65:10854-10861. Dongming m, Gaobin p, Caiyan L (2009). Isolation and Characterization of AaWRKY1, an Artemisia annua Transcription Factor that Regulates the Amorpha-4,11-diene Synthase Gene, a Key Gene of Artemisinin Biosynthesis. Plant and cell physiology. 50 (12):2146-2161. ISSN: 0032-0781 CODE: PCPHA5. Doyle JJ, Doyle JL (1990). Isolation of plant DNA from fresh tissue. Focus. 12:1315. Dudareva N, Martin D, Kish CM, Kolosova N, Gorenstein N, Faldt J, Miller B, Bohlmann J (2003). E-beta-ocimene and myrcene synthase genes of floral scent biosynthesis in snapdragon: function and expression of three terpene synthase genes of a new terpene synthase subfamily. Plant Cell 15: 1227-1241. Duke MV, Paul RN, Elsohly HN, Sturtz G, Duke SO (1994). Localization of artemisinin and artemisitene in foliar tissues of glanded and glandless biotype of Artemisia annua L. Int. J. Plant Sci. 155: 365-372. Duke SO, Vaughn KC, Croom EM, Elsohly HN (1987). Artemisinin, a constituent of annual worm wood (Artemisia annua), is a selective phytotoxin. Weed Sci. 35: 499-505. Durand-Tardif M, Broglie R, Slightom J, Tepfer D (2006). Structure and expression of Ri-TDNA from A. rhizogenes in Nicotiana tobacum. Organ and phenotypic specificity. J. Mol. Biol. 186: 557-564. Eckstein LU, Webb RJ, Van GI, East JM, Lee AG, Kimura M, O’Neill PM, Bray PG, Ward SA, Krishna S (2003). Artemisinin target the SERCA of Plasmodium falciparum. Nature 424: 957-961. Efferth T (2007). Antiplasmodial and antitumor activity of artemisinin from bench to bedside.PlantaMedica. 73: 299–309. Efferth T (2009). Artemisinin: A Versatile Weapon in Traditional Chinese Medicine. In: Ramawat KG (ed), Herbal Drugs: Ethnomedicine to Modern Medicine, Springer-Verlag Berlin Heidelberg p. 173-194. Efferth T, Davey M, Olbrich A, Rucker G, Gebhart E, Davey R (2002). Activity of drugs from traditional Chinese medicine toward sensitive and MDR1-or MDR1-overexpressing multidrug-resistant human CCRF-CEM leukemia cells. Blood Cells, Molecules, and Diseases 28, 160–168. Efferth T, Dunstan H, Sauerberry A, Miyachi H, Chitambar CR (2001). The antimalarial artesunate is also active against cancer. International Journal of Oncology 18: 767-773. Efferth T, Marschall M, Wang X, Huong SM, Hauber I, Olbrich A, Kronschnabl M, Stamminger T, Huang ES (2002). Antiviral activity of artesunate towards wild-type, recombinant, and ganciclovir-resistant human cytomegaloviruses. J. Mol. Med. 80: 233-242. Efferth T, Oesch F (2004). Oxidative stress response of tumor cells: microarraybased comparison between artemisinins and anthracyclines. Biochemical Pharmacology 68, 3–10. Efferth T, Olbrich A, Bauer R (2002). mRNA expression profiles for the response of human tumor cell lines to the antimalarial drugs artesunate, arteether, and artemether. Biochem Pharmacol. 64: 617–23. Efferth T, Romero MR, Wolf DG, Stamminger T, Marin JJG, Marschall M (2008). The antiviral activities of artemisinin and artesunate. Clinical Infectious Diseases 47: 804–811. Efferth Thomas (2006). Molecular Pharmacology and Pharmacogenomics of Artemisinin and its Derivatives in Cancer Cells. Current Drug Targets. 7 (4): 407-421. Eldomiaty MM, Almeshal IA, Elferaly FS (1991). Reversed phase HPLC determination of artemisitene in artemisinin. J. Liq. Chromatogr. 14: 23172330. Elford BC, Roborts ME, Phillipson JD, Wilson R JM (1987). Potentiation of the antimalarial activity of qinghaosu by methoxylated flavonoids. Transactions of the royal society of Tropical medicines and hygiene. 81: 434-436. Elsohly HN, Croom EM, Elferaly FS, Elsherei MM (1990). A large-scale extraction technique of artemisinin from Artemisia annua. J. Nat. Prod. 53: 1560-1564. Elsohly HN, Croom EM, Elsohly MA (1987). Analysis of the antimalarial sesquiterpine artemisinin in Artemisia annua by high performance liquid chromatography (HPLC) with post column derivatization and ultraviolet detection. Pharmacol. Res. 4: 258-260. Estruch JJ, chriqu, D, Grossmann K, Schell J, spena A (1999). The plant oncogene rol C is responsible for the release of cytokinias from glucoside conjugates. EMBO J. 10: 2889-2895. Estruch JJ, Schell J, spena A (1999). The protein encoded by the rol B plant oncogene hydrolysis indole glucosides. EMBO. J. 10:3125-3128. Eugene V, Michael YG (2003). Sequence - Evolution – Function. Computational Approaches in Comparative Genomics. ISBN-10: 1-40207-274-0. Fatokun CA, Menacio DI, Danesh D, Young ND (1992). Evidence for orthologous seed weight genes in cowpea and mungbean, based upon RFLP mapping. Genetics 132, 841–846. Fazal M, Karim BH, Khan, AB (1997). Effect of Afsanteen (Artemisia absinthium) in acute intestinal amoebiasis. Hamdard Medicus. 40: 24-27. Ferreira JF, Simon JE, Janick J (1997). Artemisia annua: botany, horticulture, pharmacology. Hort. Rev. 19: 319-371. Ferreira JFS, Charles DJ, Wood K, Janick J, Simon JE (1994). A comparison of gas chromatography and high performance liquid chromatography for artemisinin analyses. Phytochem. Anal. 5, 116-120. Ferreira JFS, Janick J (1996). Roots as an enhancing factor for the production of artemisinin in shoot cultures of Artemisia annua. Plant Cell Tissue Organ Cult. 44: 211-217. Ferreira JFS, Janick J (1996b). Roots as an enhancing factor for the production of artemisinin in shoot cultures of Artemisia annua. Plant Cell Tissue Organ Cult. 44: 211-217. Ferreira JFS, Janick J (2002). Production of artemisinin from in vitro cultures of Artemisia annua. In: Nogata T, Chizuka Y (eds), Biotechnology in agriculture and forestry, Medicinal and Aromatic Plants XII Springa-Verlag Berlin. 51: 112. Ferreira JFS, Jannick J in J. Jannik (1996). (Eds.) Progress in new crops ASHS Press Arlington: 579-8. Ferreira JFS, Laughlin JC, Delabays N, Magalhaes PM (2005). Cultivation and genetics of Artemisia annua for increased production of the anti-malarial artemisinin. Plant Gen. Resourc. 3: 206-229. Ferreira JFS, Simon JE, Janick J (1995). Developmental studies of Artemisia annua: flowering and artemisinin production under greenhouse and field conditions. Planta Medica. 61: 167–170. Ferreira JFS, Simon JE, Janick J (1995a). Developmental studies of Artemisia annua L: Flowering and artemisinin production under green house and field conditions. Planta Med. 61: 167-170. Ferreira JFS, Simon JE, Janick J (1995b). Relationship of artemisinin content of tissue cultured, greenhouse grown and field grown plants of Artemisia annua. Planta Med. 61: 351-355. Filip CW, Van N, Sofie RF, Vande C, Lies M, Alain G, Dirk I, Jan VB, Dieter LD (2006). Quantitation of artemisinin and its biosynthetic precursors in Artemisia annua L. by high performance liquid chromatographyelectrospray quadrupole time-of-flight tandem mass spectrometry. J. Chromatogr. A 1118: 180-187. Filippini F, Rasi V, Marin O, Trovato M, Downey PM, Lo Shiavo F, Terzi M (1994). A plant oncogene as a phosphatase. Nature. 379: 499-500. Filippini F, Rasi V, Marin O, Trovato M, Downey PM, Lo Shiavo F, Terzi M (1996). A plant oncogene as a phosphatase. Natrue. 379: 499-500. Flavell RB, Bennett MD, Smith JB, Smith DB (1974). Genome size and the proportion of repeated nucleotide sequence DNA in plants. Biochem. Genet. 12, 257–269. Flores H, Filner P (2007). In: Primary and secondary metabolism of plant cell culture. Neuman, K., Barz, W., Reinhard, E. (Eds.). Springer Berlin Heidelberg, New York. Pp.174-185. Flores HE, Curtis WR (1992). Approaches to understanding and manipulating the biosynthetic potential of plant roots. In: Pederson H, Mutharsan R, DiBiasio D (eds), Biochemical Engineering VII: Cellular and Reaction Engineering, Proc. New York Acad. Sci. 665: 188-209. Flores HE, Filner P (1985). Metabolic relationships of putrescine, GABA and alkaloids in cell and root cultures of Solanaceae. In: Neumann KH, Barz W, Reinhard E (eds), Primary and Secondary Metabolism of Plant Cell Cultures. Springer-Verlag, Berlin p. 174-185. Fogh J, Trempelln G (1975). In: Fogh, J. _Ed.., Human Tumor Cells in vitro. Plenum Press, New York, p. 115. Fulzele DP, Sipahimalani AT, Heble MR (1991). Tissue cultures of Artemisia annua: organogenisis and artemisinin production. Phytother. Res. 5: 149-153. Galal EE, Kandil A, Abdel-Latif M, Kheds T, Khafagy SM (1974). Cardiac pharmaco toxicological studies of Judaicin, isolated from Artemisia Judaica planta Med. 25: 88-91. Gamborg OL, Miller RA, Ojima K (1968). Transfer of T-DNA from Agrobacterium to the plant cell. Plant Physiol. 107: 1041. Garfield E (1986). Schistosomiasis: The Scourge of the Third World, Part 1. Etiology. Curr. Comm. 9(9): 3-7. Garfinkel DJ, Simpson RB, Ream LW, White FF, Gordon MP, Nester EW (2006). Genetic analysis of crown gall: five structure map of the T-DNA by site-directed mutagenesis. Cell. 27: 143-153. Gey GO, Coffman WD, Kubicek MT (1952). Tissue culture studies of the proliferative capacity of cervical carcinoma and normal epithelium. Cancer Res. 12, 264. Ghafoor A (2002). Asteraceae. In: Flora of Pakistan – Ali, S. I. and Qaiser, M. (Eds.). Missuri Botanical Press. Ghafoor A, Al-Turki TA (2000). Flora of Kingdom of Saudi Arabia. Ali, S. I. and Qaiser chaudhary, (Eds.). II (3): 182-186. Gilani, AH, Jambaz KH (1995). Preventive and curative effects of Artemisia dubia on autominophen and carbontetra chloride induced hepatotoxicity. Gen. Pharmacy. 26(2): 309-315. Giri A, Narasu M (2000). Transgenic hairy roots. Recent trends and applications. Biotechnol Adv.; 18:1–22. Graham IA, Besser K, Blumer S, Branigan CA, Czechowski T, EliasL,Guterman I, Harvey D, Isaac PG, Khan MA, LarsonTR, Li Y, PawsonT,Penfield T, Rae AM, Rathbone DA, Reid S, Ross J, Smallwood MF, Segura V, Townsend TD, Vyas, Winzer T, BowlesD (2010). The Genetic Map of Artemisia annua L. Identifies Loci affecting yield of the antimalarial drug artemisinin. Report science. 327 (5963):328-331. Green MD, Mount DL, Todd GD, Capomacchia AC (1995). Chemiluminescent detection of artemisinin novel endoperoxide analysis using luminol without hydrogen peroxide. Journal of Chromatography A 695, 237–242. Greger H, Hofer O (1980). New unsymmetrical substituted tetrahydrofurafusan lignans from Artemisia absinthian. Assignment of the relative stereochemistry by lanthanide include chemical shift. Tetrahedron 36: 3551-3558. Grønbaek K, Hother C, Jones PA (2007). Epigenetic changes in cancer. APMIS. 115(10):1039-59. Gulati A, Bharel S, Jain SK, Abdin MZ, Srivastava PS (1996). In vitro micropropagation and flowering in Artemisia annua. J. Plant Biochem. Biotechnol. 5: 31-35. Guo PQ, Yi-Wen Y, Qi-Long R (2005). Determination of artemisinin in Artemisia annua L. by reversed phase HPLC. J. Liq. Chromat. Related Tech. 28 (5): 705-712. Hajra PK, Rao RR, Singh DK, Uniyal BP (1995). Flora of India. Botanical Survey of India, Culcutta. 12: 8-47. Hamill J, Parr A, Robins R, Rhodes M (1986). Secondary product formation by cultures of Beta vulgaris and Nicotiana rustica transformed with Agrobacterium rluzogenes. Plant cell Rep. 5: 111-114. Hampel D, Mosandl A, Wust M (2005). Biosynthesis of mono- and sesquiterpenes in carrot roots and leaves (Daucus carota L.): metabolic cross talk of cytosolic mevalonate and plastidial methylerythritol phosphate pathways. Phytochem. 66: 305-311. Han JL, Liu BY, Ye HC, Wang H, Li ZQ, Li GF (2006). Effects of Overexpression of the endogenouse farnesyl diphosphate synthase on the artemisinin content in Artemisia annua L. J Integr Plant Biol. 48:482-487. Han JL, Wang H, Ye HC, Liu Y, Li ZQ, Zhang Y, Zhang YS, Yan F, Li GF (2005). High efficiency of genetic transformation and regeneration of Artemisia annua L. via Agrobacterium tumefaciens-mediated procedure. Plant Sci 168:73–80. Handa SS, Anupam S, chakraborti KK (1986). National products and plants, as liver protecting drugs. Fitoterapia. 57: 307-351. Hänsel R, Keller K, Rimpler H, Schneider G (1992). (Hrsg) Hagers Handbuch der Pharm Praxis 5.Auflage Band 4 Springer Verlag Berlin Heidelberg New York. pp 357. Haq I (1983). Medicinal plants Hamdard Foundation Press, Karachi. Pp.51-55. Hartwell LH, Hood L, Goldberg ML, Reynolds AE, Silver LM, Veres RC (2008). Genetics: from genes to genomes. Boston: McGraw-Hill Higher Education. ISBN 0-07-284846-4. Haynes RK (2006). From artemisinin to new artemisinin antimalarials: biosynthesis, extraction old and new derivatives, stereochemistry and medicinal chemistry requirements. Current Topics in Medicinal Chemistry. 6: 509–537. He XC, Zeng MY, Li GF, Liang Z (1983) Callus induction and regeneration of plantlets from Artemisia annua and changes of qinghaosu contents. Acta Bot Sin 25:87–90 Heinrich M, Robles M, West J, Artiz D, Madellano B, Rodriguz E (1998). Ethnopharmacology of Mexian Asteraceae (Ceupasitae). Ann. Rev. Pharmacal. Texical. 38: 539-565. Henry L, Tomikazu S, Narendra PS (2005). Oncologic, endocrine and metabolic treatment of cancer with artemisinin and artemisinin tagged iron-carrying compounds. Expert Opin. Ther. Targets 9(5): 995-1007. Hernandez H, Mendiola J, Torres D, Garrido N, Peres N (1990). Effect of aqeous extract of Artemisia on the invitro culture of plasmodium falciparum. Fitoterapia. 61: 540-541. Heywood VH, Humphries GJ (1977). The biology and chemistry of the Compositae. In: Heywood, B.H., Harbone, J.B., Tunner, B.L. (Eds.), Anthemideae— Systematic Review, vol. 2. Academic Press, London, New York, San Francisco, pp. 851–898. Hobbs SLA, Warkentin TD, DeLong CMO (1993) Transgene copy number can be positively or negatively associated with transgene expression. Plant Mol Biol 21:17–26. Hooker JD, Thomson T (1881). The Flora of British India. William Clowes and Sons Ltd. London 3: 321-330. Hooykaas PJJ (2004). Transofrmation of plant cell via Agrobacteirum. Plant. Mol. Bio. 13: 327-336. Hooykaas PJJ and Schilperoort A (1992). Agrobacterium and plant genetic engineering. Plant Mol. Biol. 19: 15-38.Paul J.J. Hooykaas and Ro Hostettmann K (1987). In-vitro culture for medicinal Asteraceae species for Arhizogenes transformation Agr. Bull Soc. Fib. Sc. Wat. 76: 51-63. Hsu E (2006). The history of qinghao in the Chinese material medica. Transact. Royal Soc. Trop. Med. Hyg. 100: 505-508. Ikram M, Shafi N, Mir I, Do MN, Naguyen P, Le-Quesne PW (1987). 24-ZetaEthylcholesta-7, 22-diene-3-betaol; a possibly antipyretic constituent of Artemisia absinthium. Planta Med. 53: 389. Jaziri M, Diallo B, Vanhaelen M, Homes J, Yoshimatsu K, Shimonura K (1993). Immunodetection of artemisinin in Artemisia annua cultivated in hydroponic conditions. Phytochem. 33: 821-826. Jaziri M, Diallo B, Vanhaelen M, Homes J, Yoshimatsu K, Shimonura K (1993). Immunodetection of artemisinin in Artemisia annua cultivated in hydroponic conditions. Phytochem. 33: 821-826. Jaziri M, Shimonura K, Yoshimatsu K, Fauconnier ML, Marlier M, Homes J (1995). Establishment of normal and transformed root cultures of Artemisia annua L. for artemisinin production. J. Plant Physiol. 145: 175-177. Jeyadevan JP, Bray PG, Chadwick J, Mercer AE, Byrne A, Ward SA, Park BK, Williams DP, Cosstick R, Davies J, Higson AP, Irving E, Posner GH, O’Neill PM (2004). Antimalarial and antitumor evaluation of novel C-10 nonacetal dimers of 10beta-(2-hydroxyethyl)deoxoartemisinin. Journal of Medicinal Chemistry 47, 1290–1298. Jha J, Jha TB, Mahato SB (1988). Tissue culture of Artemisia annua L.: A potential source of an antimalarial drug. CurrSci. 57:344-346. Jiao Y, Ge CM, Meng QH (2007). Dihydroartemisinin is an inhibitor of ovarian cancer cell growth. Acta Pharmacol Sinica. 28:1045-1056. John ST, Jeroen R (2004). Duplication and divergence: The Evolution of New Genes and Old Ideas. Annual Review of Genetics. 38: 615-643. DOI: 10.1146/annurev.genet.38.072902.092831. Jorsobe KL, Wullems GJ, West J, Artiz D (2003). Agrobacterium mediated transformation and its further application to plant biology. Annu. Rev. Plant Physiol. 38: 467-486. Joss H, Inge D, Caplan A, Sormann M, Schell J (1989). Genetic analysis of TDNA transcripts in nepaline crown gall. Cell. 32: 1057-1067. Jouanin L (1988). Restriction map of an agropine type Ri plasmid and its homologies with T: Plasmids. Plasmid 12: 91-102. Jun Z, Bernardo L, Erik BD, Daniel LH (2011). Copy-Number Variation: The Balance between Gene Dosage and Expression in Drosophila melanogaster. Genome Biol Evol. 3 1014-1024. doi: 10.1093/gbe/evr023. Jung, G, Tepfer D (1987). Use of genetic transformation by the T-DNA of A. rhizogenes to stimulate biomass and tropene alkaloid production in Atropa belladonna and Calystogia sepium roots grown in-vitro plant Sci. 50: 145-151. Jun-li H, Hong W, He-chun Y, Yan L, Zhen-qiu L, Yi Z, Yan-sheng Z, Fang Y, Guo-feng L (2005). High efficiency of genetic transformation and regeneration of Artemisia annua L. via Agrobacterium tumefaciens mediated procedure. Plant Science 168 (2005) 73–80. Junshen H, Pao-Chen L, Mei C (1996). Choleretic Principles of Artemisia. 12: 281. Kamada H, Okamura N, Stake M, Harada H, Shimomura K (1986). Alkaloid production by hairy root cutlrues in Atropa belladonna. Plant Cell Rep. 5: 239242. Kamchonwongpaisan S, Meshnich SR (1996). The mode of action of antimalarial artemisinin and its derivatives. Gen. Pharmacol. 27: 587-592. Kapoor RR, Chaudhary V, Bhatnagar AK (2007). Effects of arbuscularmycorrhiza and phosphorus application on artemisinin concentration in Artemisia annua L. Mycorrhiza. 17: 581–587. Karbwang J, Bangchang KN, Molunto P, Banmairuroi V, Chongpuong K (1997). Determination of artemether and its major metabolites, dihydroartemisinin, in plasma using high performance liquid chromatography with electrochemical detection. J. Chromatogr. B. Biomed. Appl., Mar, 690 (l-2), 259-265. Kaul VK, Nigam SS, Dhar KL (1976). Antimicrobial activities of the essential oils of Artemisia absinthium Linn. Artemisia vestita wall and Artemisia vulgaris L. Ind. J. Pharmacy. 38: 21-22. Khafagy SM, El-Din AA, Jakupovic J, Zdero C, Bohlmann F (1988). Glaucolidelike sesquiterpenes lactones from Artemisia judacia. Phytochem. 27: 11251128. Khafagy SM, Tosson S (1968). Crystallographic optical and chromatograph studies of judaicin, bitter principle of Artemisia judacia L. Planta Med. 16: 446-449. Kim NC, Kim SU (1992). Biosynthesis of artemisinin from 11, 12-dihydroarteannuic acid. J. Kor. Agric. Chem. Soc. Rev. 35: 106-109. Kim Y, Wyslouzil B, Weathers PJ (2002). Secondary metabolism of hairy root cultures in bioreactors. In Vitro Cell Dev. Biol. Plant 38: 1-10. Kim Y, Wyslouzil BE, Weathers PJ (2001). A comparative study of mist and bubble column reactors in the in vitro production of artemisinin. Plant Cell Rep. 20: 451-455.States: J. Nat. Prod. 47: 715-717. Kim Y, Wyslouzil BE, Weathers PJ (2001). A comparative study of mist and bubble column reactors in the in vitro production of artemisinin. Plant Cell Rep. 20: 451-455. Kiso Y, Ogasawara S, Hirotas K, Watanabe N, Oshima Y, Kano C, Hikino H (1984). Antithepato toxic principles of Artemisia capillaries bud. Planta Med. 50: 81-85. Klayman DL (1985). Qinghaosu (artemisinin) an antimalarial drug from Cluna. Science. 228: 1049-1055. Klayman DL (1989). Weeding out Malaria. Nat. Hist. 5: 18-26. Klayman DL (1993). Artemisia annua: From weed to respectable antimalarial plant. In: Kinghorn AD, Balandrin MF (eds), Human medicinal agents from plants. Am. Chem. Soc. Symp. Series. ACS, Washington DC. p. 242-255. Klayman DL, Lin AJ, Acton N, Scovill JP, Hoch JM, Milhous WK, Theodarides AD, Debek AS (1984). Isolation of artemisinin (qinghaosu) from Artemisia annua growing in the United States: J. Nat. Prod. 47: 715-717. Knight CJ, Bailey AM, Foster GD (2010). Investigating Agrobacterium-Mediated Transformation of Verticillium albo-atrum on Plant Surfaces. PLoS ONE 5(10): e13684. doi:10.1371/journal.pone.0013684 Kohler M, Haerdi W, Christen P, Veuthey JL (1997). Extraction of artemisinin and artemisinic acid from Artemisia annua L. using supercritical carbon dioxide. J Chromatogr A. 785(1-2):353-60. Koonin EV (2005). "Orthologs, paralogs, and evolutionary genomics". Annu. Rev. Genet. 39: 309–38. DOI:10.1146/annurev.genet.39.073003.114725. PMID 16285863. Kornkven A, Watson L, Estes J (1998). Phylogenetic analysis of Artemisia section Tridentata (Asteraceae) based on sequences from the internal transcribed spacers (ITS) of nuclear ribosomal DNA. Am. J. Bot. 85: 1787-1795. Kornkven A, Watson L, Estes J (1999). A molecular phylogeny of Artemisia section. Tridentatae (Asteraceae) based on chloroplast DNA restriction site variation. Syst. Bot. 24: 69-84. Kovalenko PG, Antonjuk VP, Maliuta1 SS (2004). Secondary metabolites synthesis in transformed cells of Glycyrrhiza glabra L. and Potentilla alba L. as producents of radioprotective compounds. Ukr. Bioorg. Acta. 1(2): 13-22. Krits P, Fogelman E, Ginzberg I (2007). Potato steroidal glycoalkaloid levels and the expression of key isoprenoid metabolic genes. Planta 227: 143-150. Kudakasseril GJ (1987). Effect of sterol inhibitors on the incorporation of 14 Cisopentenyl pyrophosphate into artemisinin by a cell-free system from Artemisia annua tissue cultures and plants. Planta Med. 53: 280-284. Kudakasseril GJ (1987). Effect of sterol inhibitors on the incorporation of 14 Cisopentenyl pyrophosphate into artemisinin by a cell-free system from Artemisia annua tissue cultures and plants. Planta Med 53: 280-284. Kumar S, Gupta SK, Gupta MM, Verma RK, Jain DC, Shasany AK, Darokar MP, Khanuja SPS (2002). Method for maximization of artemisinin production by the plant Artemisia annua. US Patent 6, 393, 763. Kunitake E, Shuji T, Jun-ichi S, Takashi (2011). Agrobacterium tumefaciensmediated transformation of Aspergillus aculeatus for insertional mutagenesis. AMB Express. 1:46. Lagercrantz U, Putterill J, Coupland G, Lydiate D (1996). Comparative mapping in Arabidopsis and Brassica, fine scale genome collinearity and congruence of genes controlling flowering time. Plant J. 9, 13–20. Lai H and Singh NP (1995). Selective cancer cell cytotoxicity from exposure to dihydroartemisinin and holotransferrin. CancerLett. 91: 41–6. Lai H, Singh NP (2006). Oral artemisinin prevents and delays the development of 7, 12-dimethylbenz (a) anthracene (DMBA)-induced breast cancer in the rat Cancer Lett. 233 (1): 43-8. Lange BM, Rujan T, Martin W, Croteau R (2000). Isoprenoid biosynthesis: the evolution of two ancient and distinct pathways across genomes. Proc. Natl. Acad. Sci. U. S. A. 97: 13172-13177. Lapkin AA, Plucinski PK, Cutler M ( 2006). Comparative Assessment of Technologies for Extraction of Artemisinin. Journal of Natural Products 69 (11): 1653-64. Lapkin AA, Walker A, Sullivan N, Khambay B, Mlamboa B, Chemata S (2009). Development of HPLC analytical protocols for quantification of artemisinin in biomass and extracts. Journal of Pharmaceutical and Biomedical Analysis 49: 908–15 Laughlin JC (1995). The influence of distribution of antimalarial constituents in Artemisia annua L. on time and method of harvest. Acta Hort. 390:67-73. Lee CH, Hong H, Shin J, Jung M, Shin I, Yoon J, Lee W (2000). NMR studies on novel antitumor drug candidates, deoxoartemisinin and carboxypropyldeoxoartemisinin. Biochemical and Biophysical Research Communication 274, 359–369. Lee J, Zhou HJ, Wu XH (2006). Dihydro - artemisinin downregulates vascular endothelial growth factor expression and induces apoptosis in chronic myeloid leukemia K562 cells. Cancer Chemother Pharmacol. 57: 213–20. Lee K, Geissman T (1970). Sesquiterpene lactones of Artemisia. Caustituents of Artemisia ludoviciana ssp. Mexicana, Phytochem. 9: 403-408. Lewis W, Vinay D, Zenger V (1983). Airborne and allergic pollen of North America, Baltimore, MD, Johns Hopkins Press. Li G, Guo X, Jin R, Wang Z, Jian H, Li Z (1982). Clinical studies on treatment of cerebral malaria with qinghaosu and its derivatives. J. Trad. Chin. Med. 2: 125-130. Li Y, Shan F, Wu JM, Wu GS, Ding J, Xiao D, Yang WY, Atassi G, Leonce S, Caignard DH, Renard P (2001). Novel antitumor artemisinin derivatives targeting G1 phase of the cell cycle. Bioorg Med Chem Lett. 11: 5–8. Liersch R, Soicke H, Stehr C, Tullner HU (1986). Formation of artemisinin in Artemisia annua during one vegetation period. Planta Med. 52: 387-390. Lies M, Filip CW, Van N, Yansheng Z, Darwin WR, Jacob P, Sofie RF, Vande C, Dirk I, Patrick SC, Dieter LDD, Alain G (2011). Dissection of the phytohormonal regulation of trichome formation and biosynthesis of the antimalarial compound artemisinin in Artemisia annua plants. New Phytologist. 189: 176–189. Lin AJ, Klayman DL, Hoch JM, Silverton JV, George CF (1985). Thermal rearrangement and decomposition products of artemisinin (qinghaosu). J. Org. Chem. 50: 4504-4508. Linda M, Van B (2009). Role of Viruses in Human Evolution. yearbook of physical anthropology. 46:14–46 (2003). Ling YR (1982). On the system of the genus Artemisia L. and the relationship with its allies. Bull. Bot. Res. 2: 1-60. Ling YR (1988). The Chinese Artemisia Linn. – the classification, distribution and application of Artemisia Linn. in China. Bull. Bot. Res. 8: 1-61. Ling YR (1992). The old world Artemisia Linn. (Compositae). Bull. Bot. Res. 12: 1108. Ling YR (1994a). The New World Artemisia L. In: Hind D, Beentje H (eds), Proc. Kew Internat. Compos. Confer. London p. 255-281. Ling YR (1994b). The genera of Artemisia L. and Seriphidium (Bess) Poljak. In the World Compos. Newsl. 25: 39-45. Ling YR (1995). On the floristics of Artemisia L. in the world. Bull. Bot. Res. 15: 137. Linnaeus C (1753). Species Plantarum, Ray Society London, Reprint ed. 1959. Adlard and Son, Bartholomew Press Dorking, Great Britain. Liu B, Wang H, Du Z, Li G, Zi H (2010). Metabolic engineering of artemisinin biosynthesis in Artemisia annua L. Plant Cell Rep. 30:689–694. Liu CZ, Wang X C, Zhao B, Guo C, Ye HC, Li GF (1999). Development of a nutrient mist. Bioreactor for growth of hairy roots. In vitro cell. Dev. Biol. Plant 35: 271-274. Liu CZ, Zhoua HY, Zhaoa Y (2007). An effective method for fast determination artemisinin in Artemisia annua L. by high performance liquid chromatography with evaporative light scattering detection. Anal. Chem. Acta 581: 298-302. Liu JM, Ni MY, Fan JF, Tu YY, Wu ZH, Qu YL, Chou MS (1979). Structure and reactions of artemisinin. Huaxue Xuebao 37(2): 129-143. Liu WH, Saint DA (2002). Validation of a quantitative method for realtime PCR kinetics. Biochem. Biophys. Res. Commun. 294: 347–353. Lommen WJM, Elzinga S, Verstappen FWA, Bouwmeester HJ (2007) Artemisinin and sesquiterpene precursors in dead and green leaves of Artemisia annua L. crops Planta Medica. 73: 1133–1139. Lommen WJM, Schenk E, Bouwmeester HJ, Verstappen FWA (2006). Trichome dynamics and artemisinin accumulation during development and senescence of Artemisia annua leaves. Planta Medica. 72: 336–345. Luo XD, Shen CC (1987). The chemistry, pharmacology and clinical applications of qinghaosu (artemisinin) and its derivatives. Med. Res. Rev. 7: 29-52. Luo XD, Shen CC (1987). The chemistry, pharmacology and clinical applications of qinghaosu (artemisinin) and its derivatives. Med. Res. Rev. 7: 29-52. Lydon J, Teasdale JR, Chen PK (1997). Allelopathic activity of annual wormwood (Artemisia annua) and the role of Artemisinin. Weed Sci. 45: 807-811. Mabberley DJ (1997). The plant-book, A portable dictionary of the vascular plants, 2nd ed. Cambridge University Press, Cambridge, UK. p. 876. Mackay WA, Kitto SL (1998). Factors affecting in-vitro proliferation of French Tarragon. Hort. Sci. 113(2): 282-287. Mannan A, Ahmed I, Arshad W, Asim MF , Qureshi RA, Hussain I, Mirza B (2010). Survey of artemisinin production by diverse Artemisia species in northern Pakistan. Malaria Journal. 9:310 doi:10.1186/1475-2875-9-310. Mannan A, Mirza B (2004). Comparative study of artemisinin content in different species Artemisia available in Pakistan and antibacterial assay of Artemisia vestita. International symposium on Medicinal Plants: Linkage beyond national boundaries. NARC, Islamabad, Pakistan. p. 7-9. Mannan A, Shaheen N, Arshad W, Qureshi RA, Zia M, Mirza B (2008). Hairy roots induction and Artemisinin analysis in Artemisia dubia and Artemisia indica. Afr J Biotech. 7 (18): 3288-3292. Mannan A, Tooba N S, Mirza B (2009). Factors affecting Agrobacterium tumefacienes mediated transformation of Artemisia absinthium. Pak. J. Bot., 41(6): 3239-3246. Mano Y, Nebeshima S, Matsui C, Ohkaua H (2005). Production of tropane alkaloids by hairy root cultures of scopolia Japonica. Agri. Biol. Chem. 50: 2715-2722. Marcel K, Werner H, Phiilippe C, Jean LV (1997). Supercritical fluid extraction and chromatography of artemisinin and artemisinin acid: An improved method for the analysis of Artemisia annua samples. Phytochem. Anal. 8: 223-227. Marco JA, Barbera O (1990). Natural products from the genus Artemisia L. In: studies in natural products. Atta-ur-Rehman, (eds.). Elsevier, Amsterdam. 7(A): 201-264. Maria R, Jean FB, Jean G, Jean-Pierre L, Olivier M, Elizabeth CT (1990). Evidence for Conservation of Ferritin Sequences among Plants and Animals and for a Transit Peptide in Soybean. 265 (30):18339-18344. Maria T, Diego V, Eva PG, Laura GM, Eduardo M, Ana FM, Jose LGS, Lluis M, Luis DP (2012). A role for insulator elements in the regulation of gene expression response to hypoxia. Nucleic Acids Res.40(5): 1916–1927. Martin J, Torrell M, Valles J (2001). Palynological features as a systematic marker in Artemisia s. and related genera (Asteraceae, Anthemideae): implications for sub tribe artemisinin delimitation. Plant Biol. 4: 372-378. Martinez BC, Staba J (1988). The production of artemisinin in Artemisia annua L. Tissue cultures. Adv. Cell. Cult. 6: 69-87. Martin-Tanguy J, Frary A, Ganal MW, Spivey RT (2006). Conjugated polyamines and reproductive development biochemical, molecular and physiological approqaches. Physiol Plant. 100: 675-688. Maruzzela JC, Scavandish D, Scrandis JB, Grabon G (1960). Action of adoriferaus organic chemicals and essential oils on wood destroying fungi. Plant Disease Rep. 44: 789-792. Mary V. Duke, Rex N. Paul, Hala N. Elsohly, George S, Stephen OD (1994). Localization of Artemisinin and Artemisitene in Foliar Tissues of Glanded and Glandless Biotypes of Artemisia annua L. Int. J. Plant. Sci. 155(3): 365-372. Matsushita Y, Kang W, Charlwood BV (1996). Clong and analysis of a cDNA enconding farnesyl disphosphate synthase from Artemisia annua. Genes 172: 207-209. Maurel C, Barbier, Srygoo H, Spena A, Tempe J, Guern J (1991). Single rol genes from the A. rhizogenes TL-DNA alter some of the cellular responses to auxin in Nicotiana tobacum. Plant Physiol. 97: 212-216. Maurel C, Barbier, Srygoo H, Spena A, Tempe J, Guern J (2007). Single ral genes from the A. rhizogenes TL-DNA alter some of the cellular responses to auxin in Nicotiana tobacum. Plant Physiol. 97: 212-216. Maurel HY, Cocking EC, Yarrow SA, Burnett LA (1994). In vitro propogation of Artemisia annua L. Adv. Hort. Sci. 8: 145-147. Mauro M, Trovato M, Paolis A, Gallelli A, Castantino P, Altamura M (1994). The plant oncogene rol D stimulates flowering in transgenic tobacco plants. Dev. Biol. 180:698-1\700. McVaugh R (1984). Compositae. In: Anderson WR (ed), Flora Nova-Galiciana: a descriptive account of vascular plants of western Mexico, Vol. 12. University Michigan Press, Ann Arbor. Mercke P, Bengtsson M, Bouwmeester HJ, Posthumus MA, Brodelius PE (2000). Molecular cloning, expression, and characterization of amorpha-4,11-diene synthase, a key enzyme of artemisinin biosynthesis in Artemisia annua L. Archives of Biochemistry and Biophysics 381: 173–180. Meshnich SR (1996). Is haemozoin a target for antimalarial drugs? Ann. Trop. Med. Paracitol. 90: 367-372. Meshnich SR (1996). Is haemozoin a target for antimalarial drugs? Ann. Trop. Med. Paracitol. 90: 367-372. Meshnich SR, Yang YZ, Lima V, Kuypers F, Kamchonwongpaisan S, Yuthavong Y (1993). Iron-dependent free radical generation from antimalarial agent artemisinin (qinghaosu). Antimicrob. Agents Chemother. 37: 1108-1114. Meshnick S, Taylor T, Kamchonwongpaisian S (1996). Artemisinin and the antimalarial endoperoxides: from herbal remedy to targeted chemotherapy. Microbial. Rev. 60(2): 301-315. Meyer A, Tempe J, Castantino P (2002). Hairy root, a molecular overview. Functional analysis of Agrobacterium rhizogenes T-DNA genes. In: a Stacey, WT Keen, eds., plant microbe interactions. APS Press, St. Paul. 93-139. Michael A, Burtin D, Tepfer D (1998). Effects of rol C locus from the Ri TL-DNA of A. rhizogenes on development and polymine metabolism in tobacco. GIM 90 proceddings, vol. 2. Pp.863-868. Mingfu W, Chungheon P, Qingli W, James ES (2005). Analysis of Artemisinin in Artemisia annua L. by LC-MS with Selected Ion Monitoring. J. Agric. Food Chem. 53: 7010-7013. Moore M (1979). Medicinal plants of the mountain west. Museum of New Mexico Press, Santa Fe, New Mexico p. 162. Morgan AJ, Cox PN, Turner DA, Peel E, Davey MR, Gratland KMA, Mulligan BJ (2004). Transformation of Artemisia annua using Ti plasmid vector. Plant Sci. 49:37-49. Mukanganyama S, Widersten M, Naik YS, Mannervik B, Hasler JA (2002). Inhibition of glutathione S-transferases by antimalarial drugs possible implications for circumventing anticancer drug resistance. International Journal of Cancer 97, 700–705. Murai L, Kemp JD (1988). Octopine synthase mRNA isolated from sunflower crown gall callus is havologaus to the Ti plasmid of Agrobacterium tumefaciences. Proc. Watl. Acad. Sci. USA. 79: 86-90. Murashige T, Skoog E (1962). A revised medium for rapid growth and bioassay with tobacco tissue culture. Physiol Plant. 15: 472-493. Murata M (1998). Effect of genotype and explant type in regeneration frequency of Artemisia plant in vitro. Acta Horti. 231-234. Nair MS, Acton N, Klayman DL (2001). Production of artemisinin in tissue cultures of Artemisia annua L. J. Nat. Prod. 49: 504-507. Nair MSR, Acton N, Klayman DL, Kendrick K, Basile DV, Mante S (1986). Production of artemisinin in tissue cultures of Artemisia annua. J Nat Prod. 49:504-507. Nam W, Tak J, Ryu JK, Jung M, Yook JI, Kim HJ (2007). Effects of artemisinin and its derivatives on growth inhibition andapoptosis of oral cancer cells. Head Neck. 29: 335–40. Navaratnam V, Mordi MN, Mansor SM (1997). Simultaneous determination of artesunic acid and dihydroartemisinin in blood plasma by using high performance liquid chromatography for application in clinical pharmacological studies. J. Chromatogr. B. Biomed. Appl. 692(l): 157-162. Newman JD, Marshall J, Chang M, Nowroozi F, Paradise E, Pitera D, Newman KL, Keasling JD (2006). High-level production of amorpha-4,11-diene in a two-phase partitioning bioreactor of metabolically engineered Escherichia coli. Biotechnol. Bioeng. 95: 684-691. Nilsson O, Olsson O (1997). Getting to the root: The role of the Agrobacterium rhizogenes rol genes in the formation of hairy roots. Physiol. Plant. 100: 463473. Nin A, Bennici G, Roselli D, Mariotti S, Schiff RM (1997). Agrobacterium- mediated transformation of Artemisia absinthium L. (wormwood) and production of secondary metabolites. Plant Cell Reports. 16: 725–730. Nin S, Morosi E, Schiff S, Bennici A (1996). Callus cultures of Artemisia absinthium L.: initiation, growth optimization and organogenesis. Plant Cell Tiss. Org. Cult. 45: 67-72. Nin S, Schiff S, Bermici A, Magherini R (2004). In vitro propogation of Artemisia annua L. Adv. Hort. Sci. 8: 145-147. Norma BP, Ana Maria G (1995). Artemisinin production by Artemisia annua L.transformed organ cultures. Enzyme and Microbial Technology DOI:10.1016/0141-0229:00216-2 Ohta T (1989). Role of gene duplication in evolution. Genome. 31(1):304-10. Ohta T (2000). Evolution of gene families. Gene. 23;259(1-2):45-52. Old RW, Primrose SB (1995). Principles of gene manipulations on introduction to Genetic Engineering. Blookwell Science, pp.278. Olliaro PL, Nair NK, Sathasivam K, Mansor SM, Navaratnam V (2001a). Pharmacokinetics of artesunate after single oral administration to rats. BMC Pharmacology, 1, 12. Olofsson L, Anneli L, Peter EB (2011). Trichome isolation with and without fixation using laser micro dissection and pressure catapulting followed by RNA amplification: Expression of genes of terpene metabolism in apical and sub-apical trichome cells of Artemisia annua L. Plant Science. 183: 9– 13. Olsson ME, Olofsson LM, Lindahl AL, Lundgren A, BrodeliusM,Brodelius PE (2009). Localization of enzymes of artemisinin biosynthesis to the apical cells of glandular trichomes of Artemisia annua L. Phytochemistry. 70: 1123–1128. Ooms G, Twell D, Base ME, Harry C, Hoge C, Burrell MM (2008). Developmental regulation of Ri-TL-DNA gene expression in roots, shoots and Tubers of transformed potato. Plant Mol. Biol. 6: 321-330. Oshima, Y., Kawakami, Y., KisoY, Hikino H, Yang LL, Yen KY (1984). Antihepatoxic principles of Aeginatia indica herbs. Shoyakugaku Zasshi 38: 198-200. Palazón J, Cusidó RM, Gonzalo J, Bonfill M, Morales S, Piñol MT (1998b). Relation between the amount the rolC gene product and indole alkaloid accumulation in Catharantus roseus transformed root cultures. J Plant Physiol. 153:712–8. Palazon J, Cusido RM, Raig C, dinal M T (1997). Effect of ral genes from A. rhizogenes TL-DNA on nicotine production in tobacco root cultures. J. Plant. Physiol. Biochem. 35: 155-162. Palazón J, Cusidó RM, Roig C, Piñol MT (1998a). Expression of the rolC gene and nicotine production in transgenic roots and their regenerated plants. Plant Cell Rep. 17:384–90. Palozon J, Cuswido RM, Ganzalio J, Banfill M, Marales C, Pinnal T (2003). Relation between the amount of rol C gene product and Indole alkaloid accumulation in catharanthus roseaus transformed root cultures. J. Plant Physiol. 153: 712-718. Paniego NB and AM Giuliette (1994). Artemisia annua L.: dedifferentiated and differentiated cultures. Plant Cell Tiss Organ Cul.t 36:163-168. Parmeshwari S, Ram AV (1998). HPLC-electrospray ionization mass spectrometric analysis of antimalarial drug artemisinin. Anal. Chem. 70: 3084-3087. Paterson AH, Lin Y.-R, Li Z, Schertz KF, Doebley JF, Pinson SRM, Liu SC, Stansel JW, Irvine JE (1995). Convergent domestication of cereal crops by independent mutations at corresponding genetic loci. Science 269, 1714–1718. Patrick RA, Daniel RV, Kristin KW, Pamela JW (2010). Effect of Sugars on Artemisinin Production in Artemisia annua L. Transcription and Metabolite Measurements. Molecules. 15: 2302-2318. Pellegrino AP, Joaquin DRG, Shepherd KSL (2008). In vitro culture, four medicinal Asteraceae species for A. rhizogenes transformation Agr. Prod. Post-Harvest Techniques, Biotechnology. 729: 299-301. Peres LEP, Morgante PG, Vecclu C, Kraus JE, Van Sluys MA (2001). Shoot regeneration capacity from rats and transgenic hairy roots of Artemisia annua cultivars and wild related species. Plant Cell Tiss. Org. Cult. 65: 37-44(8). Petersen NJ, Barrett DH, Bond WW, Berquist KR, Favero MS, Bender TR, Maynard JE (1976). Hepatitis B surface antigen in saliva, impetiginous lesions, and the environment in two remote Alaskan villages. Appl. Environ. Microbiol. 32(4): 572-574. Podlech D (1986). Compositae VI-Anthemidae. In: Rechinger KH (eds), Flora Iranica. Akademische, Druck-U, Verlagsanstalt Graz-Austria p.159-223. Polunin O, Staintan A (1992). Flowers of Himalya. Oxford University Press Delhi. 193-195. Portela A, Esteller M (2010). Epigenetic modifications and human disease. Nat Biotechnol. 28(10):1057-68. Posada D, Crandall KA, Holmes EC (2002). Recombination in evolutionary genomics. Annu Rev Genet. 36:75-97. Posner GH, McRiner AJ, Paik IH, Sur S, Borstnik K, Xie S, Shapiro TA, Alagbala A, Foster B (2004). Anticancer and antimalarial efficacy and safety of artemisinin-derived trioxane dimers in rodents. J Med Chem. 47: 1299– 1302. Posner GH, Park IH, Sur S, McRiner AJ, Borstnik K, Xie S, Shapiro TA (2003). Orally active, antimalarial, anticancer, artemisinin-derived trioxane dimers with high stability and efficacy. Journal of Medicinal Chemistry 46, 1060– 1065. Posner GH, Ploypradith P, Parker MH, O’Dowd H, Woo SH, Northrop J, Krasavin M, Dolan P, Kensler TW, Xie S, Shapiro TA (1999). Antimalarial, antiproliferative, and antitumor activities of artemisinin-derived, chemically robust, trioxane dimers. Journal of Medicinal Chemistry 42, 4275– 4280. Pras N, Visser JF, Batterman S, Woerdenbag HJ, Malingre TM, Lugt CB (1991). Laboratory selection of Artemisia annua L. for high artemisinin yielding types. Phytochem. Anal. 2: 80-83. Price R, Van Vugt M, Phaipun L, Luxemburger C, Simpson J, McGready R,Kuile FT, Khan A, Chongsuphajasiddhi T, White NJ, Nosten F (1999). Adverseeffects in patients with acute falciparum malaria treated with artemisinin derivatives. Am. J. Trop. Med. Hyg., 60 (4), 547-555 Prinsen E, Bercetche J, Chriqui D, Onckelen VH (1992). Pisum sativum epicotyes inoculated with A. rhizogenes agropine strains harbauring various T-DNA fragments. Morphology, histology and endogenous indale 3-acetic-acid and indale 3-actamide content. J. Plant Physial. 140: 75-83. Prinsen E, Bytabier B, Hernalsteener JP, De Greef J, Onckelen VH (1998). Emotional expression of A. tumefacienes T-DNA onco genes in Asparagus crown gall tissue. Plant all Physiol. 31: 69-75. Qin MB, Li GZ, Ye HC, Li GF (1994). Induction of hairy root from Artemisia annua with Agrobacterium Rhizogenes and its culture in vitro. Acta Bot Sin 36: 165-170. Ranashige A, Sweatlock JD, Cooks RG (1993). A rapid screening method for artemisinin and its congeners using ms / ms: search for new analogs in Artemisia annua L. J. Nat. Prod. 56: 552-563. Rao VSN, Menezes AMS, Gadelha MGT (1988). Antifertility screening of some indigenous plants of Brazil. Fitoterapia. 59: 17-20. Regis S, Grossi S, Corsolini F, Biancheri R, Filocamo M (2009). PLP1 gene duplication causes overexpression and alteration of the PLP/DM20 splicing balance in fibroblasts from Pelizaeus-Merzbacher disease patients. Biochim Biophys Acta. 1792(6):548-54. Renata F, Maria FPSM, Valéria SL, João FVS, Carlos AAA, Silvana RRM, Eliseu B, Ricardo VA, Francismar CM, Alexandre LN (2010). Size of AT(n) Insertions in Promoter Region Modulates Gmhsp17.6-L mRNA Transcript Levels. J Biomed Biotechnol. 2010: 847673. Reungpatthanapong P, Mankhetkorn S (2002). Modulation of multidrug resistance by artemisinin, artesunate and dihydroartemisinin in K562/adr and GLC/adr resistant cell lines. Biological and Pharmaceutical Bulletin 25, 1555–1561. Riggs AD, Russo VEA, Martienssen RA (1996). Epigenetic mechanisms of gene regulation. Plainview, N.Y: Cold Spring Harbor Laboratory Press. ISBN 0-87969-490-4. Rizwana AQ, Mushtaq A, Muhammad A (2002). Taxonomic study and medicinal importance of three selected species of the genus Artemisia Linn. Asian J. Plant Sci. 1(6): 712-714. Ro DK, Paradise EM, Ouellet M, Fisher KJ, Newman KL, Ndungu JM, Ho KA, Eachus RA, Ham TS, Kirby J, Chang MC, Withers ST, Shiba Y, Sarpong R, Keasling JD (2006). Production of the antimalarial drug precursor artemisinic acid in engineered yeast. Nature 440: 940-943. Robertson D, Earle ED (2000). Plant regeneration from leaf explant of Artemisia annua. Plant Cell. Rep. 5: 61-64. Robinson WS, Clayton DA, Greenman RL (1974). DNA of a Human Hepatitis B Virus Candidate. J. Virol. 14(2): 384-391. Rohmer M, Seemann M, Horbach S, Bringermeyer S, Sahm H (1996). Glyceraldehyde 3-phosphate and pyruvate as precursors of isoprenic units in an alternative non-mevalonate pathway for terpenoid biosynthesis. J. Am. Chem. Soci. 118: 2654-2566. Romero MR, Efferth T, Serrano MA, Castano B, Macias RI, Briz O, Marin JJ (2005). Effect of artemisinin/artesunate as inhibitors of hepatitis B virus production in an in vitro replication system. Antiviral. Res. 68: 75-83. Roth RJ, Acton N (1989). The isolation of sesquiterpines from Artemisia annua. J. Chem. Educ. 66: 349-350. Rowen RJ (2002). Artemisinin: From Malaria to Cancer treatment. Townsend Lett. Doct. Patients 86-88. Royal Society (2002). Genetically Modified Plants for Food Use and Human Health—An Update. Policy Document 4/02. The Royal Society, London. Ryde´n A-M, Ruyter-Spira C, Quax WJ, Osada H, Muranaka T, Kayser O, Bouwmeester H (2010). The molecular cloning of dihydroartemisinic aldehyde reductase and its implication in artemisinin biosynthesis in Artemisia annua. Planta Medica. doi: 10.1055/s-0030-1249930. Sa G, Mi M, He-Chun Y, Ben-Ye L, Guo-feng LKC (2001). Effects of ipt gene expression on the physiological and chemical characteristics of Artemisia annua L. Plant Sci. 160:691–698. Sachs J, Malaney P (2002). The economic and social burden of malaria. Nature 415: 680-685. Sadava D, Phillips T, Lin C, Kane SE (2002). Transferrin overcomes drug resistance to artemisinin in human small-cell lung carcinoma cells. Cancer Lett. 179: 151-156. Sadeghpour O, Asghari G, Ardekani MRS, Jaroszewski JW (2006). Phytochemical study of Artemisia persica boiss. and evaluation of its antiplasmodial activity. Plant. Med. 72(11): 1023-1023. Said HM (1969). Hamdard Pharmacopeae of Eastern Medicines, Institute of Health and Tibbi Research, Hamdard foundation Press, Karachi p. 361. Said HM (1982). Drug used in liver disease since ancient times. In the disease of liver: Greco-Arab Concepts. Hamdard Foundation press, Karachi p. 100-131. Said HM (1996). Medicinal Herbal MAS printers Karachi, Pakistan. 1: 34-35. Saito K, Yamazaki M, Anzai H, Yoneyama K, Murakoshi I (2001). Transgenic resistant. Atropa bella Donna using an Ri binary vector and inheritance of transgenic trait. Plant. Cell. Rep. 11: 219-224. Sambrook LF, Adams TR, Schell JS, Nester EW (1989). Agrobacterium tumefaciens mediated transformarion of Artemisia abscinthium stem explants by using kanamycin selection. Proc Natl Acad Sci USA. 85: 5536-5540. Sandrenan N, Sioufi A, Godbillon J, Netter C, Donker M, Van V (1997). Determination of artemether and its metabolite, dihydroartemisinin, in plasma by high-performance liquid chromatography and electrochemical detection in the reductive mode. Journal of chromatography B: Biomedical applications, Amsterdam. 691 (1): 145-153. Sanford KK, Earle WR, Likely GD (1948). The growth in vitro of single isolated tissue cells. J. Natl. Cancer Inst. 9, 229. Sangwan RS, Agarwal K, Luthra R, Thakur RS, Singh-Sangwan N (1993). Biotransformation of arteannuic acid into arteannuin-B and artemisinin in Artemisia annua. Phytochemistry 34: 1301–1302. Schmid G, Hofheinz W, Hart P, Duke SO (2005). Efficient and genotype independent Agrobacterium mediated transformation of Artemisia annua. J Plant Physiol. 160(10): 1253-1257 Schmid G, HofheinzW (1983). Total synthesis of qinghaosu. J Am Chem Soc. 105: 624 –625. Schmidt M, Khan A, Schmidt AM, Heinze B, Hack E, Waltenberger J, Kreienberg R (2007). A novel breast cancer cell line initially established from pleural effusion: Evolution towards a more aggressive phenotype. Int J Oncol. 30: 565-572. Schmulling T, Fladung M, Grossmann K, Schell J (1993). Hormonal content and sensitivity of transgenic tobacco and potato plants expressing single rol genes of A. rhizogenes T-DNA. Plant J. 3: 371-382. Schramek N, Wang H, Ro¨misch-Margl W, Keil B, Radykewicz T, Winzenho¨rlein B, Beerhues L, Bacher A, Rohdich F, Gershenzon J (2010). Artemisinin biosynthesis in growing plants of Artemisia annua. A 13CO2 study. Phytochemistry 71: 179–187. Seeger C, Mason WS (2000). Hepatitis B virus biology. Microbiol. Mol. Biol. Rev. 64 (1): 51-68. Sergeeva NV, Zakharova NL (1977). Qualitative and quantitative study of carotenoides in some galenical preparations. Farmatsiya 26: 34-38. Sevon N and Oksman-Caldentey, KM (2002). Agrobacterium rhizogenes-mediated transformation: root cultures as a source of alkaloids. Planta Medica. 68 (10): 859-868. Sheela Chandra (2012). Natural plant genetic engineer Agrobacterium rhizogenes: role of T-DNA in plant secondary metabolism. Biotechnol Lett. 34:407–415. DOI 10.1007/s10529-011-0785-3. J.J. Hooykaas Shen WH, Daviaud E, David C, Barbier-Brygoo H, Tupe J, Guern J (2003). High sensitivity to auxin is a common feature of hairy root. Plant Physiol. 94: 554560. Shen WH, Petit A, Guern J, Tempe J (2001). Hairy roots are more sensitive to auxin than normal roots. Proc. Natl. Acad. Sci. USA. 85:3417-3421. Shiba Y, Paradise EM, Kirby J, Ro DK, Keasling JD (2007). Engineering of the pyruvate dehydrogenase bypass in Saccharomyces cerevisiae for high-level production of isoprenoids. Metabol. Eng. 9: 160-168. Shkryl YN, Veremeichik GN, Bulgakov VP, Tchernoded GK, Mischenko NP, Fedoreyev SA (2007). Individual and combined effects of the rol A, B and C genes on anthraquinone production in Rubia cordifolia transformed calli. Biotechnol Bioeng .doi:10.1002/bit.21727published online. Singh A, Vishwakarma RA, Husain A (1988). Evaluation of Artemisia annua strains for higher artemisinin production. Planta Med. 54: 475-476. Singh NP and Lai H (2001). Selective toxicity of dihydroartemisinin and holotransferrin toward human breast cancer cells. Life Sci. 70: 49-56. Singh NP and Lai HC (2004). Artemisinin induces apoptosis in human cancer cells. Anticancer Res.24: 2277–80. Sipahimalani AT, Fulzele DP, Heble MR (1991). Rapid method for detection and determination of artemisinin by gas chromatogr. J. Chromatogr. 538: 452-455. Skehan P (1995). Assays of cell growth and cytotoxicity. In: Studzinski _Ed.., Cell Growth and Apoptosis. A Practical Approach. Oxford University Press, New York. p. 169. Skehan P, Storeng R, Scudiero D, Monks A, McMahon J, Vistica D, Warren JT, Bokesch H, Kenney S, Boyd MR (1990). New colorimetric cytotoxicity assay for anticancer- drug screening. J. Natl. Cancer. Inst. 82: 1107-1112. Skoog F, Miller CO (1997). A revised medium for rapid growth and bioassay with Artemisia tissue culture. Plant Physiol. 102:363-372. Slepetys J (1975). Biology and biochemistry of wormwood. 8. Accumulation dynamics of tannins, ascorbic acid and carotene. Liet. TSR. Mokslu. Akad. Darb Ser. C. 1: 43-48. Slightom JL, Durand-Tardif M, Jouanin L, Tepfer D (1986). Nucleotide sequence analysis of TL-DNA of Agrobacterium rhizogenes agropine type plasmid. Identification of open reading frames. J. Biol. Chem. 261: 108-121. Smith AE, Secoy DM (1988). Plants used for agricultural pest control in Western Europe before 1850. Chem. Ind. 3: 12-17 Smith TC, Weathers PJ, Cheetham RD (1997). Effect of gibberellic acid on hairy root cultures of Artemisia annua growth and artemisinin production. In Vitro Cell Dev. Biol. 33: 75-79. Sommer L, Pislarasu S, Burlibasa C (1965). Effect of chamazulene extracted from absinth on the inflammation of the periodonthium and pastoperative cauplications. Farna cia. 13: 471-474. Soule HD, Vazquez J, Long A, Albert S, Brennan M (1973). A human cell line from a pleural effusion derived from a breast carcinoma. J Natl Cancer Inst. 51 (5): 1409–1416. Souret FF, Kim Y, Wyslouzil BE, Wobbe KK, Weathers PJ (2003). Scale-up of Artemisia annua L. hairy root cultures produces complex patterns of gene expression. Biotechnol. Bioeng. 83: 653-657. Spano L, Mariotti D, Bramea C, Costantino C (2006). Morphogenesis and auxin sensitivity of transgenic tobacco with different couplements of the Ri T-DNA plant Physiol. 87: 479-483. Spano L, Mariotti D, Cardarelli M, Branca C, Costantino C (1988). Morphogenesis and auxin sensitivity of transgenic tobacco with different complements of the Ri T-DNA. Plant Physiol. 87: 479-483. Spena A, Schmulling T, Konez C, Schell JS (2002). Independent and synergistic activity of rol A, B and C loci in stimulating abnormal growth in plants. EMBO J. 6: 3891-3899. Srivastava HK (1999). Genetic diversity and enhancement for antimalarial compound in Artemisia annua. Proc. Nat. Acad. Sci. India. 69(B): 13-26. Stanton B, Gelvin (2003). Agrobacterium mediated plant transformation. The biology behind the gene-jockeying tool. Mic. Biol. Mal. Bio. Rev. 67(1): 1637. Stewart RR (1972). An annotated catalogue of the vascular plants of West Pakistan and Kashmir. Karachi. Pp. 714-721. Stiles LH, Leather GR, Chen PK (2005). Effects of two sesquiterpene lactones isolated from Artemisia annua on physiology of Lemna minor. J. Chem. Ecol. 20(4): 969-978. Stoger EA (1991). Arzneibuch der chinesischen Medizin. Deutscher ApothekerVerlag, Stuttgart, Germany. Street HE (1977). Introduction. In: plant tissue and cell culture. Street, H. E. (ed.). 2nd ed. Botanical Monogrpahs Vol.11. Blackwell Scientific Publications Oxford, London. Pp: 1-10. Stringham RW. Kenneth GL, Peter M, Greg K, István P, Christina K (2009). High performance liquid chromatographic evaluation of artemisinin, raw material in the synthesis of Artesunate and Artemether. Journal of Chromatography A, 1216: 8918–8925. Sun LY, Touraud G, Charbonnier C, Tepfer D (1999). Modification of phenotype in cicharium intybus through genetic transformation by A. rhizogenes conversion from biennial to annual flowering. Tranv. Res. 1: 14-22. Sun WC, Han JX, Yang WY, Deng DA, Yue XF (1992). Antitumor activities of 4 derivatives of artemisic acid and artemisinin B in vitro. Zhongguo Yao Li Xue Bao 13, 541–543. Sushil KSK, Gupta PS, Pratima BMM, Digvijay SAK, Gupta GK, Shasany SS (2004). High yields of artemisinin by multi-harvest of Artemisia annua crops. J. Industrial Crops and Products 19 (2004) 77–90. Sy LK, Brown GD (1998). Three sesquiterpines from Artemisia annua. Phytochem. 48: 1207-1211. Sy LK, Brown GD (2001). Deoxyarteannuin B, dihydro-deoxyarteannuin B and trans-5-hydroxy-2-isopropenyl-5-methylhex-3-en-1-ol from Artemisia annua. Phytochem. 58: 1159-1166. Sy L-K, Brown GD (2002). The mechanism of the spontaneous autoxidation of dihydroartemisinic acid. Tetrahedron 58: 897–908. Tahir M, Khan AB, Siddique MMH (1991). Effect of Artemisia absinthium L. In acute intestinal amoebicasis. In: Conference of pharmacology and symposium on Herbal drugs, 15th March, New Delhi, India. Pp.20. Tan RX, Lu H, Walfender JL, Yu TT, Zheng WF, Yang L, Gafner S, Hostettmann K (1991). Mona and sesquiterpines and antifungal constituents from Artemisia species. Planta Med. 65: 64-67. Tan RX, Lu H, Wolfender JL, Yu TT, Zheng WF, Yang L, Gafner S, Hostettmann K (1999). Mono and sesquiterpines and antifungal constituents from Artemisia species. Planta Med. 65: 64-67. Tan W, Jinjian L, Mingqing H, Yingbo L, Meiwan C, Guosheng W, Jian G, Zhangfeng Z, Zengtao X, Yuanye D, Jiajie G, Xiuping C, Yitao W (2011). Anti-cancer natural products isolated from chinese medicinal herbs. Chinese Medicine. 6:27. Tawfiq NK, Anderzson LA, Roberts MF, Phillipson JD, Bray DH, Warhurst DC (1989). Antiplasmodial activity of Artemisia annua plant cell culture. Plant Cell Rep. 8: 425-428. Taylor GR (1991). Polymerase chain reaction: basic principles and automation. In: PCR: A practical approach. (McPherson, M.J., Quikie, P., Taylor, GR.eds). IRL Press, Oxford. Pp: 1-14. Tellez MR, Canel C, Rimando AM, Duke SO (1999). Differential accumulation of isoprenoids in glanded and glandless Artemisia annua. L. Photochemistry. 52: 1035–1040. Teoh KH, Polichuk DR, Reed DW, Covello PS (2009). Molecular cloning of an aldehyde dehydrogenase implicated in artemisinin biosynthesis in Artemisia annua. Botany 87: 635–642. Teoh KH, Polichuk DR, Reed DW, Nowak G, Covello PS (2006). Artemisia annua L. (Asteraceae) trichome-specific cDNAs reveal CYP71AV1, a cytochrome P450 with a key role in the biosynthesis of the antimalarial sesquiterpene lactone artemisinin. FEBS Letters 580: 1411–1416. Thomas E, Axel S, Armin O, Erich G, Pia R, Oliver W (2003). Molecular Modes of Action of Artesunate in Tumor Cell Lines. Mol Pharmacol. 64:382–394. Torrel M, Garcia-Jacas N, Susanna A, Valles J (1999). Phylogeny in Artemisia (Asteraceae, Anthemidae) inferred from nuclear DNA (ITS) sequences. Taxon 48: 721-736. Towler MJ, Weathers PJ (2007). Evidence of artemisinin production from IPP stemming from both the mevalonate and the nonmevalonate pathways. Plant Cell Rep. 26:2129-2136. Trovato MP, Brooks B, Seki JM, Gordan MP (2001). Recent advances in fruit development and ripening: an overview. J. Exp. Bot. 53: 377. Tu YY, Ni MY, Zhong YR, Li LN, Cui SL, Zhang MQ, Wang YZ, Ji Z, Liang XT (1982). Studies on constituents of Artemisia annua. Part II. Planta Med. 44: 143-145. Tue Nguyen Michel Arnoux (2011). Semi-synthetic Artemisinin Project.RBM/UNITAID/WHO ARTEMISININ CONFERENCE 2011. HANOI, VIETNAM – 2/3rd November. Valecha N, Biswas S, Badoni V, Bhandari KS, Sati OP (1994). Antimalarial activity of Artemisia japonica, Artemisia maritima, and Artemisia nilegarica. Ind. J. Pharmacol. 26: 144-146. Valles J, Torrell M, Garnatje T, Garcia-Jacas N, Vilatersana R, Susanna A (2003). The genus Artemisia and its allies: phylogeny of the subtribe Artemisiinae (Asteraceae, Anthemideae) based on nucleotide sequences of nuclear ribosomal DNA internal transcribed spacers (ITS). Plant Biol. 5: 274284. Van der Kooy F, Verpoorte R (2011). The content of artemisinin in the Artemisia annua tea infusion. Planta Medica-Natural Products and Medicinal Plant Research. 77(15): 1754-1756. Van Oncekelon HA, Rudelsheim P, Inze D, Follin A, Messens E, Horemams S, Schell J, De Greef J (1985). Tobacco plants transformed with the Agrobacterium T-DNA gene I contain high amounts of Indale-3-acetamide. FEBS Lett. 131: 373-376. Van Quekelberghe SA, Soomro SA, Cordonnier JA, Jansen FH (2008). Optimization of an LC-MS method for the determination of artesunate and dihydroartemisinin plasma levels using liquid-liquid extraction. J Anal Toxicol. 32(2):133-9. Vanicha Vichai and Kanyawim Kirtikara (2006). Sulforhodamine B colorimetric assay for cytotoxicity screening. The Nature. Doi: 10.1030/nprot.2006.179. Vatsya B, Bhaskaran S, Thorpe TA (2002). Transformation of Artemisia with Agrobacterium tumefaciens. Plant Tiss Cult ManB6: 1144. Vergauwe A, Cammart R, Vandenberghe D, Genetello C, Inze D, Van-Montagu M, Van Den Eeckhout E (2002a). Agrobacterium tumefacienes mediated transformation of A. annua L. plants and the regeneration of transgonic plants. Plant cell Rep. 15: 929-933. Vergauwe A, Geldre EV, Inze D, Van Montagu M, Eeckhout VD (1998). Factors influencing Agrobacterium tumefaciens-mediated transformation of Artemisia annua L. Plant Cell Reports. 18: 105–110. Vergauwe A, Inze D, Van Gelder E, Van Montagu M, Van Den Eeckhout E (1996b). The use of amoxicillin and ticarcillin in combination with a betalactamase inhibitor as decontaminating agent in the Agrobacterium tumefaciens mediated transformation of Artemisia dubia L. J. Biotechnol. 52:89-95. Vergauwe A, Ronny C, Dirk, Christiane G, Dirk I, Marc VM, Elfride VE (1996). Agrobacterium tumefacienes-mediated transformation of Artemisia annua L. and regeneration of transgenic plants. Plant Cell Reports. 15:929-933. Vergauwe A, Van Geldre E, Inze D, Van Montagu M, Vanden Eeckhout E (2005). Factors influencing A. tumefacienes mediated transformation of Artemisia annua L. Plant Cell Rep. 18: 105-110. Vergauwe A, Van Geldre E, Inze D, Vanden Eeckhout E (1998). Factors influencing A. tumefacienes mediated transformation of Artemisia annua L. Plant Cell Rep. 18: 105-110. Vogeli U, Chappell J (1988). Induction of sesquiterpene cyclase and supression of squalene synthetase activities in plant cell cultures treated with fungal elicitor. Plant Physiol. 88: 1291-1296. Vonwiller SC, Haynes RK, King GK, Wang H (1993). An improved method for the isolation of qinghaosu (artemisinic) acid from Artemisia annua. Planta Med. 59: 562-563. Vormefelde SB, W Poser (2000). Hyperforin in extracts of St. John‘s Wort (Hypericum perforatum) for depression. Archives of Interna medicine. [Print] September 160(16): 2548-2549. Waleerat B, Chalermpol K, Masahiro M, Kanyaratt S (2010). Overexpression of farnesyl pyrophosphate synthase (FPS) gene affected artemisinin content and growth of Artemisia annua L. Plant Cell Tiss Organ Cult. 103:255– 265.DOI 10.1007/s11240-010-9775.S. Wallaart TE, Bougtsson J, Hille L, Poppingats NCA (2001). Hormonal content and sensitivity of transgenic tobacco and potato plants expressing single rol genes of A. rhizogenes T-DNA. Plant J. 212:460-468. Wallaart TE, Bouwmeester HJ, Hille J, Poppinga L, Maijers NCA (2001). Amorpha-4, 11-diene synthase: cloning and functional expression of a key enzyme in the biosynthetic pathway of the novel antimalarial drug artemisinin. Planta 212: 460–465. Wallaart TE, Pras N, Beekman AC, Quax WJ (2000). Seasonal variation of artemisinin and its biosynthetic precursors in plants of Artemisia annua of different geographical origin: proof for the existence of chemotypes. Planta Medica. 66: 57–62. Wallaart TE, Pras N, Quax, WJ (1999). Seasonal variations of artemisinin and its biosynthetic precursors in tetraploid Artemisia annua plants compared with the diploid wild-type. Planta Med. 65 723–728. Wallaart TE, Pras N, Quaz WJ (1999). Isolation and identification of dihydroartenisinic acid hydroperoxide from Artenisia annua: A novel biosyntheotic precursor of artenisinin. J. Nat. Prood. 62: 1160-1162. Wang CW (1961). The forest of China, with a survey of grassland and desert vegetations. In: Harvard Univ. Maria Moor Cabot Foundation 5 Harvard University, Cambridge, MA. p. 171-187. Wang H, Ge L, Ye HC, CHong K, Liu BY, Li GF (2004). Studies on the effects of fpf1 gene on Artemisia annua flowering time and on the linkage between flowering and artemisinin biosynthesis. Planta Med. 70:347-352. Wang H, Liu Y, Chong K, Liu BY, Ye HC, Li ZQ, Yan F, Li GF (2007). Earlier flowering induced by over-expression of CO gene does not accompany increase of artemisinin biosynthesis in Artemisia annua. Plant Biol. 9:442446. Wang H, Ma C, Li Z, Ma L, Wang H, Ye H, Xu G, Liu B (2010). Effects of exogenous methyl jasmonate on artemisinin biosynthesis and secondary metabolism in Artemisia annua L. Industrial crops and products. 31: 214218. Wang WJ, Tan XR (2002). Artemisinin production in Artemisia dubia hairy root cultures with improved growsth by altering the nitrogen source in the medium. Biotechnol. Letters. 24: 1153-1156. Warren KS (1987). Schistosomiasis: past, present and future. Mem. Inst. Oswaldo Cruz, Vol. 82. Intern. Symp. Schistosomiasis p. 25-29. Watson LE (2002). Molecular phylogeny of subtribe Artemisunal (Asteraceae), including Artemisia and its allied and segregate genera. Biomed Cultural Evolutionary Biology2 (17). Weathers P, Smith T, Hemmavanh D, Follansbee E, Ryan J, Cheetham R (1996). Production of the antimalarial, artemisinin, by transformed roots of Artemisia annua. Acta Hort. 426: 157-163. Weathers PJ, Bunk G, Mccoy MC (2005). The effect of phytohormones on growth and artemisinin production in Artemisia annua hairy roots. In Vitro Cell. Dev. Biol. Plant. 41: 47-53 Weathers PJ, Cheetham RD, Follansbee E, Theoharides K (1994). Artemisinin production by transformed roots of Artemisia annua. Biotechnol. Lett. 16: 1281-1286. Weathers PJ, DeJesus-Gonzalez L, Kim YJ, Souret FF, Towler MJ (2004). Alteration of biomass and artemisinin production in Artemisia annua hairy roots by media sterilization method and sugars. Plant Cell Rep. 23: 414-418. Weathers PJ, Elkholy S, Wobbe KK (2006). Artemisinin: the biosynthetic pathway and its regulation in Artemisia annua, a terpenoid-rich species. In Vitro Cell Dev-Pl 42: 309–317. Weathers PJ, Hemmavanh D, Walcerz DB, Cheetham RD, Smith TC (1997). Interactive effects of nitrate and phosphate salts, sucrose, and inoculum culture age on growth and sesquiterpene production in Artemisia annua hairy root cultures. In Vitro Cell Dev. Biol. Plant, 33: 306-312. Wei L, Guangqum G, Guochang Z (2000). Agrobacterium mediated transformation: State of the art and future prospect. Chinese Sci. Bulletin, 45(17): 1537-1546. Welch BL (1947) The generalization of "Student's" problem when several different population variances are involved. Biometrika 34 (1–2): 28–35. White FF, Nester EW (2003). Relationship of plasmids responsible for hairy root and crown-gall tumorigeniety. J. Bacterial. 144: 710-720. White FF, Taylor BH, Huffman GA, Gardon MP, Nester EW (2008). Molecular and genetic analysis of the transferred regions of the root inducing plasmid of A. rhizogenes. J. Bacterial. 164: 33-34. White FF, Taylor BH, Huffman GA, Gordon MP, Nester EW (1985). Molecular and genetic analysis of the transferred DNA regions of the root-inducing plasmid of Agrobacterium rhizogenes. J. Bacteriol. 164: 33-44. WHO (2008 update). WHO forecast, In: Artepal, the portal of information and orientation on malaria and its treatments with ACT, Bangkok. WHO (2008a). World Malaria Report. WHO (2008b). Schistosomiasis. Fact Sheet. WHO (2008c). Hepatitis B. Fact Sheet No 204. WHO (2008d). Cancer. Fact Sheet No 297. Widmer V, Handloser D, Reich E (2007). Quantitative HPTLC analysis of Artemisinin in dried Artemisia annua. a practical approach. J. Liq. Chromatog. Related Tech. 30: 2209-2219. Willimtizer L, Sanchez-Serrano J, Buschfels E, Scholl J (2006). DNA from Agrobacteirum rhizogenes genes is transferred to and expressed in axenic hairy root plant tissues. Mol. Gen. Genet. 186: 16-22. Wilson P (1997). The pilot-scale cultivation of transformed roots. In: Doran P (ed), Hairy roots: Culture and applications. Amsterdam: Hardwood Academy Publishers p. 179-190. Winefield C, Lewis D, Arathoon S, Deroles S (1999). Alteration of petunia plant from through the interaction of the ral C gene from A. rhizogenes. Mal. Breed. 5: 543-551. Winstanley PA, Ward SA, Snow RW (2002). Clinical status and implications of antimalarial drug resistance. Microb. Infect. 4: 157-164. Woerdenbag HJ, Lugt CB, Pras N (1990). Artemisia annua L., a source of novel anti-malarial drug. Pharmaceut. Weekblat. 12: 169-181. Woerdenbag HJ, Moskal TA, Pras N, Malingre TM (1993). Cytotoxicity of artemisinin-related endoperoxides to Ehrlich ascites tumor cells. Journal of Natural Products 56, 849–856. Woerdenbag HJ, Pras N, Bos R, Visser JF, Hendriks H, Malingre TM (1991). Analysis of artemisinin and related sesquiterpenoids from Artemisia annua L. by combined gas chromatography / mass spectrometry. Phytochem. Anal. 2: 215-219. Woerdenbag HJ, Pras N, Chan NG, Bang BT, Bos R, Von UW, Van YP, Boi NV, Batterman S, Lugt CB (1994). Artemisinin related sesquiterpenes and essential oil in Artemisia annua during a vegetation period in Vietnam. Planta Med. 60: 272-275. World Health Organization (WHO). (1998). Malaria Medicines and Supplies Service (MMSS). Jeneva, 51,4. World Health Organization (WHO). (2004). Malaria Medicines and Supplies Service (MMSS). Retrieved September 21, 2006, from http://www.rollbackmalaria.org. Wright CW, Linley PA , Brun R , Wittlin S, Hsu E (2010). Ancient Chinese Methods Are Remarkably Effective for the Preparation of Artemisinin-Rich Extracts of Qing Hao with Potent Antimalarial Activity. Molecules. 15:804812; doi: 10.3390/molecules15020804. Wu JM, Shan F, Wu GS, Li Y, Ding J, Xiao D, Han JX, Atassi G, Leonce S, Caignard DH, Renard P (2001). Synthesis and cytotoxicity of artemisinin derivatives containing cyanoarylmethyl group. Eur J Med Chem. 36: 469–479. Wu XH, Zhou HJ, Lee J (2006). Dihydro - artemisinin inhibits angiogenesis induced by multiple myeloma RPMI8226 cells under hypoxic conditions via downregulation of vascular endothelial growth factor expression and suppression of vascular endothelial growth factor secretion. Anticancer Drugs.17: 839–48. Xu XJ, Zhu DZ, Wei-Shan Z (1986). Total synthesis of arteannuin and deoxyarteannuin. Tetrahedron 42: 819-828. Xu ZQ, Jia JF (1996). Callus formation from protoplast of Artemisia sphaerocephala-krasch and same factors influencing proto-plast division. Plant cell Tiss. Org. Cult. 44:129-184. Yamachika E, Habte T, Oda D (2004). Artemisinin: an alternative treatment for oral squamous cell carcinoma. Anticancer Res. 24:2153–2160. Yamada T, Palm CJ, Brooks B, Kosuge T (1985). Nucleotide sequences of the Pseudomonas savastanoi indoleacetic acid genes show homology with Agrobacterium tumefaciens T-DNA. Proc Natl Acad Sci U S A. 82(19):6522–6526. Yan L, Huisheng L, Fei P (2009). Solubility of Artemisinin in Seven Different Pure Solvents from (283.15 to 323.15) K. J. Chem. Eng. Data. 54: 762–764. Yang RY, Zeng XM, Lu YY, Lu WJ, Feng LL, Yang XQ, Zeng QP (2009). Senescent leaves of Artemisia annua are one of the most active organs for Overexpression of artemisinin biosynthesis responsible genes upon burst of singlet oxygen. Planta Medica. 76: 734–742. Yarnell E, Abascal K (2004a). Botanical prevention and treatment of malaria, Part 1: Herbal Mosquito repellant. Altern. Complement Ther. 10(4): 206-210. Yarnell E, Abascal K (2004b). Botanical prevention and treatment of malaria, Part 2: Selected Botanicals. Altern. Complement Ther. 10(50): 277-284. Yi Hong Wang (2008). How effective is T-DNA insertional mutagenesis in Arabidopsis? J Biochem Tech. 1(1):11-20. ISSN: 0974-2328. Yoshioka H, Yamada N, Doke N (1999). cDNA Cloning of sesquiterpene cyclase and squalene synthase, and expression of the genes in potato tuber infected with Phytophthora infestans. Plant Cell Physiol. 40: 993-998. Yun KW, Kil BS (1992). Assessment of allelopathic potential in Artemisia princeps var orientalis residues. J. Chem. Ecol. 18(11): 1933-1940. Zafar MH, Hamdard ME, Hameed A (1990). Screening of Artemisia absinthium for antimalarial effects of plasmodium berghei in mice, a premilinary report. J. Ethnopharmacy. 30: 223-226. Zambryski P, Tempe J, Schell J (1989). Transfer and function of T-DNA genes from Agrobacterium Ti and ri-plasmids in plant Cell. 56: 193-201. Zambryski PC (1992). Chronicles from the Agrobacterium plant cell DNA transfer story. Annu. Rev. Plant Physiol. Plant, Mol. Biol. 43: 465-490. Zeng Q, Qui F, Yuan L (2008). Production of artemisinin by genetically-modified microbes. Biotechnol. Lett. 30: 581-592. Zhang F, Gosser DK, Meshnich SR (1992). Hemin-catalysed decomposition of artemisinin (qinghaosu). Biochem. Pharmacol. 43: 1805-1809. Zhang L, Jing F, Li F, Li M, Wang Y, Wang G, Sun X, Tang K (2009). Development of transgenic Artemisia annua (Chinese wormwood) plants with an enhanced content of artemisinin, an effective anti-malarial drug, by hairpin-RNA-mediated gene silencing. Biotechnol Appl Biochem. 52(3):199-207. Zhang Y, Teoh KH, Reed DW, Maes L, Goossens A, Olson DJH, Ross ARS, Covello PS (2008). The molecular cloning of artemisinic aldehyde D11 (13) reductase and its role in glandular trichome-dependent biosynthesis of artemisinin in Artemisia annua. Journal of Biological Chemistry 283: 21501– 21508. Zhao KC, Liu CX, Liang XT, Mingguang Y, Song ZY (1986). Development of radioimmunoassay for determination of artesunate, a derivative of antimalarial qinghaosu. Proc. CAMS PUMC. 1: 213-217. Zhao SS, Zeng M (1985). Spectrometric high pressure liquid chromatography (HPLC) studies on the analysis of qinghaosu. Plant Med. J. Med. Plant Res. 3: 233-237. Zheng M, Li L, Chen S (1983). Chemical transformation of qinghaosu, a peroxidic antimalarial. Tetrahedron 36: 2941-2946. Zhou WS, Xu XX (1988). The structures, reactions and synthesis of arteannuin (qinghaosu) and related compounds. In: Rahman A (ed), Studies in natural products chemistry. Elsevier, New York p. 495-527. Zhou Z, Huang Y, Xie G, Sun X, Wary Y, Fu L, Jian H, Guo X, Li G (1988). HPLC with polarographic detection of artemisinin and its derivatives and application of the method to the pharmacokinetic study of artemether, J. Liq. Chromatogr. 11: 1117-1137. Zhu J, Oger PM, Schrammeijer B, Hookaas P JJ, Farrand SK, Winans SC (2000). The bases of crown gall tumorigenesis. J. Bacterial. 182: 3885-3895. Zia M, Abdul M, Chaudhary MF (2007). Effect of growth regulators and amino acids on artemisinin production in the callus of Artemisia absinthium. Pak J Bot. 39:799-805. Zuker A, Trifera T, Scovel G, Ovadis M, Skklarman E, Itzhaki H, Vainstein A (2001). rol C transgenic carnation with improved horticultural traits: quantitative and qualitative analysis of greenhouse grown plants. J. Am. Soc. Hort. Sci. 126:13-18. Zupan JR, Zambryski PC (1995). Transfer of T-DNA from Agrobacterium to the plant cell. Plant Physiol. 107: 1041. Zupan JR, Zambryski PC (2004). Transfer of T-DNA from Agrobacteirum to the plant cell. Plant Physiol. 107: 1041. Publications Publications 1. Bushra Hafeez Kiani, Naila Safdar, Abdul Mannan and Bushra Mirza (2012). Comparative Artemisinin analysis in Artemisia dubia transformed with two different Agrobacteria harbouring rol ABC genes. POJ. 5(4):386391. 2. Ali M, Kiani BH, Mannan A, Ismail T and Mirza B (2012). Enhanced production of artemisinin by hairy root cultures of Artemisia dubia. Journal of Plants Medicinal Research. 6(9): 1619-1622. 3. Bushra Hafeez Kiani, Bushra Mirza and Guy Cameron Barker (2012). Artemisinin production in Artemisia annua and Artemisia dubia following transformation with the rol ABC genes and elucidation of the sites of its synthesis. Plant Physiology. (Submitted) 4. Bushra Hafeez Kiani, Nazif Ullah, Ihsan Ul Haq and Bushra Mirza (2012). Comparative pharmacological evaluation of rol ABC genes transformed Artemisia dubia with untransformed plants. BMC Biotechnology. (Submitted) 5. Bushra Hafeez Kiani, John Suberu, Guy Cameron Barker and Bushra Mirza (2013). Development of efficient miniprep transformation methods for Artemisia annua using Agrobacterium tumefaciens and Agrobacterium rhizogenes. Plant Breeding. (Submitted) 6. Bushra Hafeez Kiani, Muhammad Tahir Waheed, John Suberu, Guy Cameron Barker and Bushra Mirza (2012). Anticancer activities of transformed and untransformed Artemisia annua and determination of functional compounds. Phytomedicine. (Submitted) Appendices APPENDIX-I YMB medium K2HPO4 0.5mg/l MgSO4 2g/l NaCl 0.1g/l Mannitol 10g/l Yeast extracts 10g/l pH 7.0 APPENDIX-II MYA Medium Yeast extracts Cas-amino acid 5g/l 0.5mg/l Mannitol 8g/l NH4 (SO4)2 2g/l NaCl 5g/l pH 6.6 APPENDIX-III Composition of Murashige and Skoog (1962) medium S. No Constituents Formula Conc. in stock solutions g/l Volume of stock/l of medium (ml) Macronutrients 1 Potassium nitrate KNO3 38 2 Ammonium nitrate NH4NO3 33 3 Calcium chloride CaCl2.2H2O 8.8 4 Magnesium sulphate MgSO4.7H2O 7.4 5 Potassium Phosphate KH2PO4 3.4 50 Micronutrients 6 Manganese sulphate MnSO4. H2O 4.4 7 Zinc sulphate ZnSO4.H2O 1.72 8 Boric acid H3BO3 1.24 9 Potassium iodide KI 1.67 10 Sodium molybdate Na2MoO4.2H2O 0.05 11 Copper Sulphate CuSO4.2H2O 0.01 12 Cobalt chloride CoCl2.6H2O 0.005 5 Iron Source 13 Sodium EDTA Na2EDTA.2H2O 7.46 14 Ferrous sulphate FeSO4.7H2O 5.56 5 Organic Supplements (Vitamins) 15 Myo-inositol 20 16 Glycine 0.4 17 Nicotinic acid 0.1 5 0.1 18 Pyridoxine-HCl 19 Thiamine-HCl 0.1 Carbon Source 20 30 g /l of Sucrose was used in MS medium. Appendix IV: For B5 Stock solution preparation S No Ingredient Name Company Name Stock Name Concentration g/L USE 1 Ammonium sulfate A-4915 (Sigma) 1.01M (NH4)2 SO4 134 1ml/L 2 Boric acid B-0252 (Sigma) 0.048M H3BO3 3 1ml/L 3 Calcium chloride dehydrate Cobalt(II) chloride hexahydrate Copper(II) sulfate pentahydrate Magnesium sulfate heptahydrate Manganese(II) sulfate monohydrate Potassium iodide C-2536 (Sigma) 1.02M CaCl2.2H2O 0.1mM CoCl2.6H2O 0.1mM CuSO4.5H2O 1.01M MgSO4.7H2O 0.059M MnSO4.H2O 150 1ml/L 0.025 1ml/L 0.025 1ml/L 250 1ml/L 10 1ml/L 4 5 6 7 8 9 Ethylenediaminetet raacetic acid disodium salt dehydrate 202185 (Sigma) 209198 (Sigma) M-7774 (Sigma) M7899 (Sigma) 207969 (Sigma) 4.5mM KI 0.750 1ml/L E-5134 (Sigma) 0.11M Na2EDTA.2H2O 37.3 1ml/L 10 Sodium molybdate dehydrate Sodium phosphate monobasic dehydrate Zinc sulfate heptahydrate M-1651 (Sigma) 13 14 11 12 15 16 17 1.03mM Na2MoO4.2H2O 1.08M NaH2PO4.2H2O 0.25 1ml/L 150 1ml/L Z-0251 (Sigma) 6.9mM ZnSO4.7H2O 2 1ml/L myo-Inositol I-3011 (Sigma) 9.91 10ml/L Thiamine hydrochloride Pyridoxine hydrochloride Nicotinic acid T-1270 (Sigma) 0.055M MyoInositol 0.03M ThiaminHCl 4.86mM Pyridoxine-HCl 8mM Nicotinic Acid 0.1M FeSO4.7H2O 2.47M KNO3 10 1ml/L 1 1ml/L 1 1ml/L 27.8 10ml/L 249.7 10ml/L Sucrose 30 30g/L 71505 (Sigma) P-8666 (Sigma) N-0765 (Sigma) F-7002 (Sigma) 18 Ferrous sulphate heptahydrate Potassium nitrate 19 Sucrose S391(Phytotechnolog y Lab) P8291 (Sigma) APPENDIX-V Southern Blot Solutions Denaturing Solution (1 liter) 0.5 N NaOH 20.0 g NaOH 1.5 M NaCl 87.7 g NaCl ddH2O to final volume Neutralizing Solution (1 liter) 1 M Tris 121.1 g Tris Base 3 M NaCl 175.4 g NaCl  fill to approximately 80% volume with ddH2O  pH to 7.0 with concentrated HCl (≈60 ml/liter)  bring to volume with ddH2O  Prehyb & Hybridization Solution (500 ml) 50% Formamide 250 ml Formamide 3X SSC 75 ml 20X SSC 1X Denhardt‘s Solution 5 ml 100X Denhardt‘s 20 μg/ml salmon sperm DNA 1 ml 10 mg/ml 5% Dextran Sulfate 25 g Dextran Sulfate 2% SDS 100 ml 10% SDS Distilled water APPENDIX-VI Contigs of A.annua and A.dubia showed 80-85% homology S.No. Contigs 1 Contig5332:0-407 2 Contig5575:0-732 3 Contig9512:0-1058 4 Contig9144:0-754 5 Contig951:0-251 6 Contig10524:0-757 7 Contig10536:0-983 8 Contig1181:0-4663 9 Contig1260:39-212 10 Contig1320:0-1308 11 Contig1746:0-2322 12 Contig2006:0-880 13 Contig2028:0-994 14 Contig2799:0-245 15 Contig2903:0-2391 16 Contig231:0-650 17 Contig2859:34-864 18 Contig2799:0-245 19 Contig3836:0-1711 20 Contig1972:29-1001 21 Contig231:0-650 22 Contig2622:14-797 23 Contig3934:0-1618 24 Contig981:16-763 25 Contig3437:0-1709 26 Contig4685:0-615 27 Contig5168:6-688 28 Contig7008:0-1433 29 Contig7150:0-1455 30 Contig7416:0-1030 31 Contig10057:0-852 32 Contig1972:29-1001 33 Contig2454:22-1644 34 Contig2799:0-245 35 Contig3437:0-1709 36 Contig7008:0-1433 37 Contig7150:0-1455 38 Contig1907:54-728 39 Contig231:0-650 40 Contig4749:0-1057 41 Contig8881:0-761 42 Contig2698:0-454 43 Contig4685:0-615 44 Contig478:36-283 45 Contig7150:0-1455 46 Contig7015:0-1455 47 Contig2006:0-880 48 Contig2288:0-726 49 Contig2454:22-1644 50 Contig2622:14-797 51 Contig290:0-475 52 Contig318:0-498 53 Contig4168:0-828 54 Contig4704:0-1313 55 Contig5025:0-1950 56 Contig5482:0-902 57 Contig583:0-1445 58 Contig6601:0-889 59 Contig762:0-855 60 Contig8568:0-1291 61 Contig4971:0-638 62 Contig5008:0-2908 63 Contig10081:0-821 64 Contig10271:0-621 65 Contig5821:0-949 66 Contig9460:0-1091 67 Contig9658:53-1094 68 Contig5624:0-2559 69 Contig7019:0-704 70 Contig489:21-1070 71 Contig7255:2-1153 72 Contig7356:0-819 73 Contig3512:0-616 74 Contig10271:0-621 75 Contig10867:0-1043 76 Contig7251:0-1059 77 Contig2139:0-1197 78 Contig359:18-539 79 Contig462:0-927 80 Contig1845:0-590 81 Contig3395:0-424 82 Contig6427:0-816 83 Contig6578:0-492 84 Contig7546:0-614