DSpace Institution DSpace Repository http://dspace.org

Animal Production and Technology Thesis and Dissertations

2019-12-24 EFFECTS OF FERTILIZER AND SOIL TYPES ON DIFFERENT CULTIVARS OF BRACHIARIA SPECIES BIOMASS YIELD AND QUALITY AND FARMERS’ PERCEPTION IN WEST GOJAM ZONE,

.Beyadglign, Hunegnaw http://hdl.handle.net/123456789/10065 Downloaded from DSpace Repository, DSpace Institution's institutional repository

BAHIR DAR UNIVERSITY COLLEGE OF AGRICULTURE AND ENVIRONMENTAL SCIENCES SCHOOL OF ANIMAL SCIENCE AND VETERINARY MEDICINE DEPARTMENT OF ANIMAL PRODUCTION AND TECHNOLOGY

GRADUATE PROGRAM MSC. IN ANIMAL PRODUCTION

EFFECTS OF FERTILIZER AND SOIL TYPES ON DIFFERENT CULTIVARS OF BRACHIARIA SPECIES BIOMASS YIELD AND QUALITY AND FARMERS’ PERCEPTION IN WEST GOJAM ZONE, ETHIOPIA

MSc. Thesis

By:

Beyadglign Hunegnaw Chekol

NOVEMBER, 2019 Bahir Dar, Ethiopia Advisers Yeshambel Mekuriaw (PhD) (Major adviser) Bimrew Asmare (PhD) (Co-adviser)

BAHIR DAR UNIVERSITY COLLEGE OF AGRICULTURE AND ENVIRONMENTAL SCIENCES SCHOOL OF ANIMAL SCIENCE AND VETERINARY MEDICINE DEPARTMENT OF ANIMAL PRODUCTION AND TECHNOLOGY GRADUATE PROGRAM: MSC. IN ANIMAL PRODUCTION

EFFECTS OF FERTILIZER AND SOIL TYPES ON DIFFERENT CULTIVARS OF BRACHIARIA SPECIES BIOMASS YIELD AND QUALITY AND FARMERS’ PERCEPTION IN WEST GOJAM ZONE, ETHIOPIA

MSc. Thesis

SUBMITTED TO THE GRADUATE PROGRAM IN THE PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER SCIENCES IN ANIMAL PRODUCTION

By

Beyadglign Hunegnaw Chekol

NOVEMBER, 2019

Bahir Dar, Ethiopia

ii

DECLARATION

This is to certify that this thesis “Effects Of Fertilizer And Soil Types On Different Cultivars Of Brachiaria Species Biomass Yield And Quality And Farmers’ Perception In West Gojam Zone, Ethiopia, Submitted in partial fulfillment of the requirements for the award of the degree of Master of Science in Animal Production to the Graduate Program of College of Agriculture and Environmental Sciences, Bahir Dar University by Mr. Beyadglign Hunegnaw Chekol (ID:BDU-1018226PR) is an authentic work carried out by him under our guidance. The matter embodied in this project work has not been submitted earlier for the award of any degree or diploma to the best of our knowledge and belief.

Name of the Student

Beyadglign Hunegnaw Chekol

Signature ______date ______

Name of the Advisers

1) Yeshambel Mekuriaw (PhD) (Major adviser)

Signature ______date______

2) Bimrew Asmare (PhD) (Co-adviser)

Signature______date______

iii

THESIS APPROVAL SHEET

As member of the Board of Examiners of the Master of Sciences (M.Sc.) thesis open defense examination, we have read and evaluated this thesis prepared by Mr. Beyadglign Hunegnaw Chekol entitled Effects Of Fertilizer And Soil Types On Different Cultivars Of Brachiaria Species Biomass Yield And Quality And Farmers’ Perception In West Gojam Zone, Ethiopia, We here by certify that, the thesis is accepted for fulfilling the requirements for the award of the degree of Master of Sciences (M.Sc.) in Animal Production.

Board of Examiners

______

Name of External Examiner Signature Date

______

Name of Internal Examiner Signature Date

______

Name of Chairman Signature Date

iv

ACKNOWLEDGMENTS

I would like to express my deepest intelligence of gratefulness, endless praises and thanks to the gigantic “GOD” and St. “Mary” for commerce his and her to get this far and for making all these possible and making me strong, healthy life, when I get adversity cursed. I wish to express my deepest gratitude and sincere appreciation to my Advisers: Dr. Yeshambel Mekuriaw and Dr. Bimrew Asmare for their constant supervision, valuable suggestion, scholastic guidance, continuous inspiration, constructive comments and immense help in field work and writing this thesis, starting from research title selection experimental procedure, data collection and final thesis write up.

Institutionally first special thanks go to Bahir Dar University College of Agriculture and Environmental Sciences School of Animal Science and Veterinary Medicine (Animal Production and Technology department) for providing me the experiment site at North Mecha . Secondly, special thanks go to Amhara Agricultural Research Institute (Andassa livestock Research Center) for granting me the study and facilitating the research process through research funding and settling the research budgets timely. Finally further special thanks go to Holeta Agricultural Research Center, Animal Nutrition Laboratory and Adet Agricultural Research Center for timely analysis of forage samples and soil chemical compositions respectively.

Personally first special thanks goes to Mulugeta Walelegn for his great contribution and follow up of lab technicians during lab analysis at Holeta Agricultural Research Center and Dr. Wubetie Adinew for sharing Brachiaria grass cultivars as root splits. I would like to express my deepest respect to all farmers which were participate at cultivars selection.

Finally, I would like to express my deepest respect to all my staff members especially the feed and nutrition case team (Adebabay Adane, Mekonnen Tilahun, Shigdaf Mekuriaw, Yehenaw Ageje, and Yhones Amsalu) for their unlimited assistance during the experimental period. My friends: Bereket Fekedie, Assemu Tesfa, Wondimenh Mekonnen, Eyasu Zeleke, Birhanu Demeke, Lissanework Mola, Wondimagene Mengesha and Birhan Kassa gave much help during my study period and for this, I extend my specials for them. Lastly, I gratefully acknowledged to my beloved parents and other relatives and well-wishers for the countless blessings, spiritual support, and sacrifice throughout my academic life.

v

LIST OF ACRONYMS AND ABBREVIATIONS

AARC Adet agricultural research center ADF Acid detergent fiber ADL Acid detergent lignin ALRC Andassa livestock research center ALT Altitude ARMA Amhara regional metrology agency CIAT International center for tropical agriculture CIMMYT International maize and wheat improvement center CP Crude protein CPY Crude protein yield CSA Central statistical agency CUFRST Cultivars, fertilizer and soil type interaction DAP Die ammonium phosphate DE Digestible energy DM Dry matter DMY Dry matter yield FAO Food and agriculture organization GB Gross benefit GDP National gross domestic product HARC Holeta agricultural research center HD Harvesting date ICIPE International center of insect physiology and ecology IGAD Inter-governmental Authority development IVOMD In vitro organic matter digestibility IVDMD In vitro dry matter digestibility Livestock and irrigation value chains for Ethiopian LIVES smallholders LLPP Leaf length per plant LSR Leaf to stem ratio LRPP Length of root per plant MRR Marginal rate of return

vi

ACRONYMS AND ABBREVIATIONS (Continued) ME Metabolizable energy MJ Mega joule MoANR Ministry of agriculture and natural resources of Ethiopia NB Net benefit NDF Neutral detergent Fiber NKP Nitrogen potassium phosphate NLPP Number of leaves per plant NPS Nitrogen phosphorus and sulfur NRC National research council NRPP Number of roots per plant PBA Partial budget analysis PH Plant height NTPP Number of tiller per plant TVC Total variable cost WF With fertilizer WOF Without fertilizer

vii

ABSTRACT

EFFECTS OF FERTILIZER AND SOIL TYPES ON DIFFERENT CULTIVARS OF BRACHIARIA SPECIES BIOMASS YIELD AND QUALITY AND FARMERS’ PERCEPTION IN WEST GOJAM ZONE, ETHIOPIA Beyadglign Hunegnaw1,2, Yeshambel Mekuriaw1 and Bimrew Asmare1,2Andassa Livestock Research Center, Amhara Regional Agricultural Research Institute (ARARI), P O Box 527, Bahir Dar, Ethiopia , 1Department of Animal production and technology, Bahir Dar University, P O Box 5501, Bahir Dar, Ethiopia Brachiaria is a high nutritive value, productive and adaptive to low rainfall and acidic soil fodder grass, which is native to tropical Africa. Yet its productivity is depends on soil type, fertilizer application, type of cultivars, agronomic practices and environmental conditions. This field experiment was conducted at black and red soils with objectives to identify and select best adaptive, productive and nutritionally good Brachiaria grass cultivars based on soil type, fertilizer effect, farmers’ perception, and partial budget analysis. A factorial arrangement of three factors (two fertilizer levels, five cultivars and two soil types) with randomized complete block design containing three replications. The Brachiaria grass cultivars used in the study were: mutica, Mulato-I, Mulato-II, Marandu and La Liberated. The root splits of cultivars were planted on well-prepared land with spacing between blocks and plots was 1m while between plants and rows was 0.5m. All biological data were collected from 10 plants grown in the two middle rows and farmers’ perception was collected before harvesting age. All Samples were harvested at 90 days of age, weighed, dried and then ground to pass l mm sieve for forage nutritive value analysis. All biological data were subjected to a general linear model analysis of variance procedures of SAS version 9.0, while farmers’ perception was captured by pair wise preference ranking. The result showed that the interaction between fertilizer, soil type and cultivars had a significant (p<0.001) effect on all morphological characteristics, dry matter yield and nutritive value of Brachiaria grasses. The application of fertilizer had a significant effect on the morphological characteristics, dry matter yield and nutritive value of Brachiaria grass cultivars at both soil types as compared to without fertilizer. All morphological characteristics and most of the nutritive value parameters except crude protein were better at black soil than red soil. In most morphological characteristics and dry matter yield, mutica cultivar had better yield at both soil types followed by hybrid Mulato-II and hybrid

viii

Mulato-I. Brachiaria Marandu, hybrid Mulato-II, and hybrid Mulato-I had better crude protein content, but Marandu had low dry matter yield. In contrast, Brachiaria mutica has the lowest nutritive value with a high fiber content of all experimental cultivars. Based on the assessment of farmers' criteria, hybrid Mulato-II, hybrid Mulato-I and mutica cultivars were selected in first, second and third order rank respectively. Moreover the higher net benefit was recorded from mutica (53447.99 and 44776) followed by hybrid Mulato-II (48648.47and 34241) and hybrid Mulato- I (28086.89 and 2939) in at black and red soil respectively with fertilizer application. Based on the overall evaluation, hybrid Mulato-II, mutica and hybrid Mulato-I cultivars were selected as adaptive and better production performance to fulfill the forage quantity and nutritive value demands in the study areas. Hence, these Brachiaria cultivars are recommended alternatively for wider adaptation and on-farm evaluation in study areas, similar soil types and agro-ecologies with fertilizer application. Keywords: assessment of farmers’, Brachiaria grass, forage economics, Plant parts and soil

ix

TABLE OF CONTENTS

Contents Pages

DECLARATION iii

THESIS APPROVAL SHEET iv

ACKNOWLEDGMENTS v

LIST OF ACRONYMS AND ABBREVIATIONS vi

ABSTRACT viii

TABLE OF CONTENTS xi

LIST OF TABLES xv

LIST OF FIGURES xvi

LIST OF APPENDIX TABLES xvii

LIST OF FIGURES IN THE APPENDIX xviii Chapter 1. INTRODUCTION 1 1.1. Background and Justification 1 1.2. Statement of the Problems 3 1.3. Objectives of the study 4 1.3.1. General objective 4 1.3.2. Specific objectives 4

Chapter 2. LITERATURE REVIEW 5 2.1. Overview of livestock feed resources in Ethiopia 5 2.1.1. Natural pasture 5 2.1.2. Crop residues 7 2.1.3. Improved forage 8 2.1.4. Agro-industrial by-products 11 2.1.5. Other feed resources 12 2.2. Overview of Brachiaria grass 12 2.2.1. Historical background of Brachiaria grass 12 2.2.2. The Biology and Agronomy of Brachiaria 13 2.3. Brachiaria grass in Ethiopia 14

xi

TABLE OF CONTENTS (continued)

2.4. Descriptions of specific Brachiaria grass cultivars 15 2.4.1. Mulato-II hybrid 15 2.4.2. Brachiaria mutica 15 2.4.3. Brachiaria brizantha 16 2.4.4. Mulato-I hybrid 17 2.4.5. Brachiaria decumbens 17 2.5. Effect of fertilizer, cultivars and soil type on plant morphological characteristics 18 2.5.1. Plant height 18 2.5.2. Number of tiller per plant 20 2.5.3. Number and length of leaves per plant 21 2.5.4. Leaf to stem ratio 22 2.5.5. Number and length of roots per plant 23 2.6. Effect of fertilizer, cultivars and soil type on forage dry matter yield 26 2.7. Effect of fertilizer, cultivars and soil type on forage quality 28 2.7.1. Dry matter content 28 2.7.2. Ash and Organic matter 28 2.7.3. Crude protein content and crude protein yield 29 2.7.4. Neutral detergent fiber 31 2.7.5. Acid detergent fiber 32 2.7.6. Acid detergent lignin 33 2.7.7. In vitro dry matter digestibility 34 2.7.8. In vitro organic matter digestibility 35 2.7.9. Metabolizable energy 35 2.8. Farmers’ perception on the varietes or cultivars selection 38

Chapter 3. MATERIALS AND METHODS 39 3.1. Description of the study areas 39 3.2. Soil sampling and chemical composition analysis 40 3.3. Experimental land preparation 42 3.4. Treatments and Experimental design 42 3.5. Data collection 43 3.5.1. Plant height 44

xii

TABLE OF CONTENTS (continued)

3.5.2. Number of tillers, leaves and leaf length per plant 44 3.5.3. Number and length of roots per plant 44 3.5.4. Leaf to stem ratio and forage yield 45 3.6. Forage quality analysis. 45 3.7. Farmer's perception on experimental Brachiaria grass cultivars 46 3.8. Partial budget analysis 47 3.9. Methods of Data Analysis 48

CHAPTER 4. RESULTS AND DISCUSSION 49 4.1. Effects of fertilizer and soil type on morphological characteristics and dry matter yield of Brachiaria grass cultivars 49 4.1.1. Plant height 49 4.1.2. Number of tillers per plant 51 4.1.3. Number of leaves per plant 53 4.1.4. Leaf length per plant 55 4.1.5. Number of roots per plant 57 4.1.6. Length root per plant 58 4.1.7. Leaf to stem ratio 59 4.1.8. Dry matter yield ton per hectare 60 4.2. Effects of fertilizer, cultivars and soil type on the quality of Brachiaria grasses 67 4.2.1. Dry matter content 67 4.2.2. Ash and Organic matter 68 4.2.3. Neutral detergent fiber 70 4.2.4. Acid detergent fiber 71 4.2.5. Acid detergent lignin 73 4.2.6. Crude protein 77 4.2.7. Crude protein yield 79 4.2.8. In vitro dry matter digestibility (IVDMD) 80 4.3.9. In vitro organic matter digestibility (IVOMD) 81 4.2.10. Metabolizable energy 82 4.3. Farmers’ perception and cultivar selection of Brachiaria grass cultivars 86 4.4. Partial Budget Analysis 88

xiii

TABLE OF CONTENTS (continued)

4.5. Correlation among morphology, dry matter yield and quality parameters of Brachiaria grass cultivars 92 Chapter 5. CONCLUSION AND RECOMMENDATIONS 95 5.1. Conclusion 95 5.2. Recommendations 96 Chapter 6. REFERENCES 97 Chapter 7. APPENDICES 112 7.1. Summary on Analysis of Variance of all experimental cultivars 112 7.2. List of Figures 119

xiv

LIST OF TABLES

Tables Pages 2.1 Average nutritive value of crop residues available in Ethiopia 8 2. 2 Avarage nutritive value some improved forages available in Ethiopia (%) 10 2.3 Average nutritive value of agro-industrial by-products available in Ethiopia 11 2.4 Summary on morphological characteristics of Brachiaria grass species by different authors 25 2.5 Summary on dry matter yield of Brachiaria grass species by different authors 27 2.6 Summary on Quantity and Quality Brachiaria grass species 36 2.7 Summary on fiber contents and digestibility of Brachiaria grass species by different authors 37 3.1 Soil sample chemical composition of experiment sites before planting 42 4.1 Effect of fertilizer, cultivars and soil type on morphological characteristics and dry matter yield of Brachiaria grasses 65 4.2 Effects of fertilizer, soil type and different cultivars on dry matter content, ash, organic matter and fiber contents of Brachiaria grass cultivars 75 4.3 Effect of fertilizer, soil type and cultivars on crude protein, crude protein yield, digestibility, and metabolizable energy Brachiaria grasses 84 4.4 Pair-wise ranking matrix of selection criteria at black soil 87 4.5 Pair-wise ranking matrix of selection criteria at red soil 87 4.6 Farmers’ preference ranking of Brachiaria grass cultivars 88 4.7 Partial budget analysis Brachiaria grass cultivars grown at black soil with and without fertilizer 90 4.8 Partial budget analysis Brachiaria grass cultivars grown at red soil with and without fertilizer 91 4.8 Correlation among on morphological parameters, dry matter yield and quality of Brachiaria grass cultivars 94

xv

LIST OF FIGURES

Figure Page 3.1 Map of study 39

xvi

LIST OF APPENDIX TABLES

Appendix Pages 7.1.1 Summary on Analysis of Variance for morphology and dry matter yield of Brachiaria grass cultivars 112 7.1.2 Summary on Analysis of Variance for chemical composition and digestibility of Brachiaria grass cultivars 113 7.1.3 Summary ANOVA on main and interaction effects on morphology and DMY of Brachiaria grass 114 7.1.4 Summary ANOVA on main and interaction effects on chemical composition and digestibility of Brachiaria grass 115 7.1.5 Pair-wise ranking of Brachiaria grass cultivars by participant farmers’ at black soil (N= 20) 116 7.1.6 Pair-wise ranking of Brachiaria grass cultivars by participant farmers’ at red soil (N=16) 117 7.1.7 Summary of on survival rate of experimental Brachiaria grass cultivars 118

xvii

LIST OF FIGURES IN THE APPENDIX

Figures Pages 7.2.1 Monthly rainfall data of experimental districts 119 7.2.2 Maximum and minimum temperature of experimental districts 119 7.2.3 Pictures of experimental Brachiaria cultivars 120

xviii

Chapter 1. INTRODUCTION

1.1. Background and Justification

Ethiopia has the largest livestock inventories in Africa having about 60.39 million cattle, 31.3 million sheep, 32.74 million goats, 11.32 million equines, 1.42 million camels and 56.06 million poultries (CSA, 2018) with livestock ownership currently contributing to the livelihoods of an estimated 80 percent of the rural population. Although variations exist among data sources, livestock contributes 15–17% of national gross domestic product (GDP), 35–47.7% of agricultural GDP and 37–87% of the household incomes (IGAD, 2011; G/Mariam et al., 2013). It also contributes 15% of export earnings and 30% of agricultural employment (Behnke, 2010). Moreover, livestock production contributes to improve the nutritional status and income gain of the people by providing meat, milk, eggs, cheese, butter, etc. and commodities, such as live animals, hides and skins for home use and export, and avert risks in times of crop failures (CSA, 2018). The livestock sector plays a significant social and economic role as a strategic tool to offset the prevailing social and economic consequences of poverty. However, it was not possible to bridge the gap between the ever- increasing demand for animal products and the level of production because of lower productivity per head of livestock (IGAD, 2011). Among the major factors responsible for the lower productivity of livestock, feed shortage both in quality and in quantity is adversely affecting livestock production (Alemayehu Mengistu et al., 2017).

The major ruminant livestock feed resources in Ethiopia are green fodder from the natural pasture (grazing) and crop residues which are poor in nutritive values (Alemayehu Mengistu et al., 2017; CSA, 2018). The natural pasture of our accounts for about 25% of total landmass (Ulfina Gelmesa et al., 2013) and it contributes about 55.96 % of the feed resources for ruminants (CSA, 208). The major limitation of natural pasture, as a source of feed for ruminant livestock is their low production of dry matter yield (Ulfina Gelmesa et al., 2013). This due to the seasonal fluctuation of rainfall and poor grazing land management, soil nutrients, conversion of grazing lands into croplands, as a result of the increased human population (Endale Yadessa et al., 2016). Besides, the available grazing lands are also overgrazed, almost one grass type and unproductive. This due to continuous heavy grazing and mismanagement of grazing lands (Abule Ebro, 2015) leading to low dry matter yield, crude protein (<5%) and metabolizable energy which results critical shortage

1

of animal feed, below the maintenance requirement of livestock (Endale Yadessa et al., 2016; Tessema Zewdu et al., 2010). Most improved tropical forage species had better forage yield and nutritional nutritive value than the naturally occurring grasses in the pasture (Alemayehu Mengistu, 2002) and a very small amount of improved forage (only 0.32 %) contribution was used as animal feed in Ethiopia (CSA, 2018).

To mitigate this shortage of quantity and quality feed, introduction and evaluation of adaptable and high yielding forage crops that enable producers to get large amounts to fodder biomass with limited land and resources are essential in the country. This technology can be also employed in the existing natural grazing land to make it more productive in terms of better botanical composition, yield and nutritive values. The improved forage crops lead to more conserved forage like hay to mitigate the feed shortage gap facing the country. So the problem of feed shortage can be addressed through identification, evaluation and promotion of forage species with better forage yield and nutritive value, which are also adapted to drought and low soil fertility (Ghimire et al., 2015). Brachiaria is a high nutritive value, productive and adaptive to low rainfall and acidic soil fodder grass, which is native to tropical Africa (Cameron, 2008). Its adaptation to the various geographical areas, the grass may be able to address the lack of year-round supply of fodder in Ethiopia. Therefore, the promising candidate improved grass could be suitable for the existing production system on the tropical ecology and adaptable for climate change are Brachiaria grass species.

Grasses in the genus Brachiaria have the advantage over those like Napier grass and Rhodes grass in other genera. These are adaptation to drought and low fertility soils, ability to sequester carbon through large root systems, increase nitrogen use efficiency through biological nitrification inhibition and arrest greenhouse gas emissions (Subbarao et al., 2009, Arango et al., 2014; Moreta et al., 2014; Rao et al., 2014). Most of the Brachiaria genus are perennial grasses, erect, semi-erect decumbent, dispersed and/or stoloniferous growth habit are strongly associated with good adaptation and production. They are adapted to a wide range of soil types, from Oxisols and Ultisols (low-fertility acid soils) to Alfisoils and Moll sols (high-fertility neutral soils) and perform much better on acid soils than other grasses such as Panicum species (Rodrigues et al., 2014). It also requires well-drained and deep soils (Njarui et al., 2015). Brachiaria grass grows well in different agro-ecological zones of tropical Africa but it performs best in sub-humid and humid environments where annual rainfalls surpass 700 mm and mean temperature exceeds 19 0c, pH ranges from 4.2 2

to 8 and hence the Brachiaria grass performs poorly 1800 m above the sea level (Larry, 2013; Vendramini et al., 2011). The productivity of the different Brachiaria grass species could be distinctly different and influenced by the area of origin, including temperature, light intensity, total rainfall, soil type, fertilizer level and by stage of maturity (Huhtanen et al., 2006). Brachiaria grass species can produce a lot of nutritious biomass on average 10 to 19.5 tons of dry matter per year per hectare when supplementing fertilizer and can also produce 6.08 tons of dry matter in the dry season (Jank et al., 2014). Crude protein content was 9 to 20% depending on soil fertility, type of cultivars, levels fertilizer application and management systems as reported by Cook et al. (2005).

The inclusion of farmers in the variety selection and assess perceptions’ enables them to select the best variety/cultivar of the new forage according to their experiences (Mutimura and Everson, 2012) on the Brachiaria grass cultivars. Nkongolo et al. (2008) also reported that participatory variety selection (PVS) helps the farmers to select better technologies by comparison to the different Brachiaria grass cultivars to forage used by farmers. Mostly researchers selected best cultivars/varieties based on morphological characteristics, agronomic data, forage yield and nutritive value of chemical composition results. This type is necessary but not sufficient for variety selection under farmers' management conditions as reported by (Misgnaw Walie et al., 2016) on cowpea varieties in northwestern, Ethiopia.

1.2. Statement of the Problems

The livestock feed resources available in the smallholder mixed production system are inadequate in quantity and low in quality mainly due to lack of suitable forages adapted to environmental conditions of the and possibly suitable to introduce into the natural pasture for improvement. Napier grass, the most widely grown fodder for the cut-and-carry in the mixed livestock production system of which currently reported to as affected by pests/smut diseases (ALRC, 2017; Mulaa et al., 2013). Rhodes grass (Chloris gayana L.), one of the cultivated pastures has a narrow genetic base and has limited ecological adaptation (HARC, 2004; CASCAPE, 2015). Crop residues and natural pasture hay are poor in nutritive values, their crude protein, minerals and vitamins are generally low, below the minimum requirement for animal production (Ulfina Galmessa et al., 2013). Recently very few studies have been carried out on the agronomic performance of Brachiaria mutica (Mimila Zemene, 2018), Brachiaria hybrid Mulato II (Negasu Yibarkew, 2018) and 3

brizantha ecotypes (Wubetie Adnew et al., 2018b). There was limited comprehensive and adequately information made in the area of evaluation and adaptation of the forage value of Brachiaria grass cultivars in Ethiopia and in the study areas with regarding of black and red soil types. Morphological characteristics, nutritive value and forage yield of fodder crops are affected by plant species, fertilizer, environmental factors (rainfall, temperature, altitude and soil type) (Cook et al., 2005). The inclusion farmers' on variety selection or perception on the Brachiaria grass cultivars in Ethiopia had limited information. Thus, participating farmers on variety selection or farmers' perception is required to enhance future wider adoption of these specific varieties /species. As to current knowledge of researchers almost all of farmers used fertilizer on forage production, they did have low information on the benefit of using fertilizer in terms of forage production and nutritionally good nutritive value. Therefore to overcome this problem understanding of plant morphology, forage nutritive value, forage yield and farmers' perception about grass information because of fertilizer application and soil type effect on Brachiaria species are principal to include it into livestock feeding system through cultivated forage production. Hence, this study was initiated to identify and select adaptable and productive Brachiaria grass cultivars in rain fed conditions under the effects of fertilizer and soil type in support with the assessment of farmers’ perception and economic evaluation.

1.3. Objectives of the study

1.3.1. General objective

 To identify and select adaptable, productive and nutritionally better Brachiaria grass cultivars based on soil type, fertilizer, farmers' perception and economic evaluation

1.3.2. Specific objectives

 To determine the effects of fertilizer, soil type and cultivars of Brachiaria grass species on forage biomass yiled and nutritional quality  To asses farmers’ perception of adaptability and production performance of Brachiaria grass cultivars  To asses economic feasibility fertilizer application on Brachiaria grass cultivars

4

Chapter 2. LITERATURE REVIEW

2.1. Overview of livestock feed resources in Ethiopia

Ruminant livestock feed resources in Ethiopia are classified as green fodder (grazing), crop residue, improved feed, hay, industrial by-products, and other feeds. Green fodder is simply pasture grasses; crop residue includes harvested by-products (straw, Stover, and hulls of cereals and pulses, etc.), improved feed is like alfalfa hay includes any type of grass, clover, etc. cut and dried as fodder. Finally industrial by-products like oilseed cakes (rapeseed cake, noug seed cake, sunflower cake, etc.), bran, and brewery residue (CSA, .2017). According to the information collected on feed usage experience of holders in the rural areas of the country, green fodder (grazing) is the major type of feed (about 55.96 %) followed by crop residue that is about 30.12% contribution for livestock feed in Ethiopia. Hay and agro- industrial by-products were used as animal feeds that comprise about 6.55 and 1.61 percent of the total feeds respectively. Moreover, a very small amount of improved feed (only 0.32 %) was used as animal feed and other types of feed that accounted for about 5.44 percent were also used in the country (CSA, 2018). Non-conventional feed source generally refer to all those feeds that have not been traditionally used for feeding livestock and are not commercially used in the production of livestock feeds. Non-conventional feeds such as vegetable refusals, sugarcane leaves, Enset leaves and fish offal used as animal feed (EndaleYadessa et al., 2016). The livestock feed resources in Ethiopia are discussed in detail in the following sections.

2.1.1. Natural pasture

Natural pasture is a predominant feed source for livestock in Ethiopia, but because of the expansion of crop farming over grazing land, overgrazing and degradation, its livestock feed contribution has become decline (Alemayehu Mengistu, 2006a; CSA, 2018). The hay harvested from the existing grazing land is characterized as low in biomass (1.5-2.0 ton dry matter ha/year) and poor in nutrients, particularly fermentable energy and degradable protein. The possible reason for its low nutritive values could be less botanical composition (poor grass-legume mixture) which in turn arise from poor management of the pasture land such as lack of or irregular fertilizer application, improper grazing management and effects of climate variability and climate change and not regularly practicing the fertilizer program.

5

As a result of this, most of the grazing lands of low lands are invaded by invasive alien species, which are plants, animals, pathogens and other organisms that are non-native to an ecosystem. In Ethiopia, natural grazing land is characterized as poor in forage diversity, biomass yield and nutritive values as consequence it demands to be intervened by introducing improved forages to make it productive pasture (Alemayehu Mengistu, 2006a).

In Ethiopia, researches on cultivated pasture and forage-crop species in the mixed crop- livestock production system had initiated in the late 1960s (Alemayehu Mengistu, 2006a). However, the adoption of cultivated pastures and forage crops in the country is not significant; less than one percent (CSA, 2018).The experience in forage research and development activities in Ethiopia shows less attention was given to develop suitable and productive forages that can be grown on natural pasture (Alemayehu Mengistu, 2006a).Yet some efforts were made on natural pasture improvement programs in Ethiopia for past years such as fencing, weeding and fertilizer application activities that have been conducted by different actors (government and non-government organizations) on the natural grazing lands to improve its status in few areas of the country. The application of nitrogen fertilizer can be improved the overall productivity of pasture as reported by (Ashagre Abate, 2008) in terms of species composition, biomass yield and nutritional values. Fertilizer of the pasture plots up to the level of 69 kg/ha improved the DMY to 36.07% over the non- fertilized plot. Harvesting pasture forages at 60 days provided the highest feeding values as measured by feed nutritive value parameters as reported by same author.

In most parts of country, farmers’ were utilize natural pasture as free grazing or in situ grazing, but in some parts they have traditional rules (by laws) how to use natural pasture especially in rainy season. During this periods they protected pasture from free grazing used as from of cut and carry system in fresh way and some are making hay for dry seasons are ways of utilizing natural pasture (Firew Tegegne and Getnet Assefa, 2010). In addition to fencing, weeding management, application of manure or fertilizer, controlled free grazing, the natural pasture productivity in our country was improved by introducing of adaptable and high-yielding cultivated forage crops with better nutritional values (Alemayehu Mengistu, 2006a). Generally grazing lands, which were the main source of natural pasture are being deteriorated due to different reasons such as high population pressure, land degradation and conversion of grazing lands into arable lands. Therefore, the production of improved fodder using different strategies like over sowing, fertilizer both organic and 6

inorganic in the natural pasture is mandatory to satisfy feed and nutrient demand of animals if better production and productivity are needed (Shimelis and Temesgen, 2016).

2.1.2. Crop residues

Crop residues are the fibrous by-products which result from the cultivation of cereals, pulses, oil plants, roots and tubers and represent an important feed resource (Yayneshet Tesfay, 2010).They are important in fulfilling feed gaps during periods of acute shortage of other feed resources .Crop-residues results from the expansion of food crops are increasing in total production from time to time and representing a large proportion of feed resources in mixed crop-livestock systems of the nation.. From the total feed biomass contribution viewpoint, the huge amount of crop residues produced could be used for livestock feeding in Ethiopia (30.12%) (CSA 2018). However, most of the cereal crop residues are of inherently low in nutritive value unless corrective management strategies are followed to exploit the feed potential of these products. Crop residues had low palatability and digestibility that leads to poor intake, particularly when fed as the sole roughage. The production of crop residues is also seasonal, available in very large quantities just after harvest and less available thereafter (Solomon Bogale et al., 2009).

Crop residues are vary greatly in chemical composition and digestibility depending on varietal differences and agronomic practices. The feeding value of crop residues is also limited by their poor voluntary intakes and low digestibility. The crude protein content of crop residues ranges between 2.4 and 7% and their IVDMD ranges between 34 and 52% Cereal straws have a mean CP, NDF, and IVDMD values of 4.5, 79.4 and 51.1%, respectively compared to pulse straws, which have a mean CP, NDF and IVDMD values of 7, 62.9 and 63.5%, respectively. Straws of oil crops have CP and NDF values of 5.4 and 66.4%, respectively (Seyoum Bediye et al., 2007). In most parts of country utilization of crop residues are after main crop harvest at filed, some conservation for next dry and wet seasons without any improvement methods as feed and also used for, fuel, construction, and bedding purposes (Firew Tegegne and Getnet Assefa, 2010). So this call up, improving nutritive value, intake and digestibility by methods like urea treatment, EM treatment, mixing of improved forages during feeding etc. are important strategies in order to maximized livestock production and productivity in small holder farmers’. Nutritive value of some crop residues available in Ethiopia are listed below (Table 2.1).

7

Table 2. 1 Average nutritive value of crop residues available in Ethiopia

Feed DM OM ADF NDF ADL CP IVDM ME ( P (%) Ca ( (%) (%) (%) (%) (%) (%) D (%) Mcal/kg) % ) Finger millet Stover 89.7 89.9 40.9 69.5 4.0 4.1 55.5 2.0 0.32 0.60 Maize Stover 91.1 92.5 47.4 70.7 5.6 4.6 58.0 2.1 0.12 0.13 Sorghum Stover 91.2 88.6 39.5 73.6 6.1 4.7 59.5 2.0 0.33 0.55 Teff straw 92.2 92.5 44.4 81.2 5.2 3.6 - - 0.11 0.30 Wheat straw 91.4 90.3 51.9 81.1 6.5 6.1 - - 0.11 Barley straw 91.1 92.4 48.3 73.9 6.2 2.4 53.5 2.0 0.13 0.44 Oats straw 91.9 91.3 36.5 58.3 4.2 7.7 62.9 2.3 0.28 0.23 Faba bean straw 92.7 90.0 56.5 67.4 11.5 - - - - - Field pea straw 87.4 93.9 - - - 4.6 - - - - Chickpea straw 91.2 90.4 42.3 55.9 9.7 5.4 0.05 1.71 Grass pea straw 90.8 95.6 11.3 58.1 0.9 - - - - - Lentil straw 91.7 92.0 44.1 64.4 9.3 8.2 55.00 2.0 - - Linseed straw 92.5 93.3 52.6 64.6 14.0 5.2 48.8 1.8 - -

(Source: Seyoum Bediye, et al., 2007, cited by Firew Tegegne and Getnet Assefa, 2010; Ethiopian Feed Database available online (http://www.vslp.org/ETHFeed/default.asp)

2.1.3. Improved forage

Forages are play variable role in different livestock production systems. Even though improved forages are important as adjuncts to crop residues and natural pastures, may be used to fill the feed gaps during periods of inadequate crop residues and natural pasture supply. Despite the improved forage production strategies had promoted in the country since in the 1940s, the contribution of improved forage as livestock, feed resource is minimal (0.32%) in Ethiopia (CSA, 2018).Utilizing improved forage varieties has several advantages like improving animal nutrition resulting in higher producing livestock, also it helps the livestock producers to conserve forage for the dry season, not only this; it also compliments crop production by maintaining soil fertility through nitrogen fixation. While grazing depletes the fertility of the land, forage growing improves soil health (Etsubdink Tekalign, 2014). The use of improved forages would reduce pressure on natural pasture, improve soil

8

fertility and erosion of marginal lands, improve carbon sequestration to mitigate climate change, support system substantially and enhance natural assets and system reliance (Etsubdink Tekalign, 2014).

The integration of improved forage species into the farming system is a promising option and /or strategy in solving chronic feed shortages in the country (Alemayehu Mengistu and Getnet Assefa, 2012). However, the adoption of improved forages at farmers' level has remained very low due to shortage of forage seed, the reluctance of most smallholder farmers (Alemayehu Mengistu and Getnet Assefa, 2012) and the lack of well-organized extension services. Similarly Abebe Mekoya et al. (2008) reported that shortage of land in the mixed crop-livestock production system, technical problems such as planting and managing the seedlings, insect damage and low interest of farmers were some of the reported reasons for poor adoption of improved forage production. Improved production and utilization practices of improved forages are also very important besides the introduction and popularization of improved forages by small-scale farmers.

Over the past four decades, several forages had tested in different agro-ecological zones, and considerable efforts have made to test the adaptability of different species of pasture and forage crops under varying agro-ecological conditions of Ethiopia. As a result, quite several useful forages had selected for different zones. Improved forage crops have been grown and used in government ranches, state farms, farmers' demonstration plots, and dairy and fattening areas. Forage crops are commonly grown for feeding dairy cattle with oats and vetch mixtures, fodder beet, elephant grass mixed with Siratro and desmodium, Rhodes/Lucerne mixture, phalaris/Trifolium mixture, hedgerows of sesbania, Leucaena and tree-Lucerne being the most common. In suitable areas, yields of oat–vetch mixtures are commonly 8–12 DM ton/ha. Yields of improved pasture and forage grasses and legumes range from 6–8 and 3–5 DM ton/ha respectively and for tree legumes 10–12 DM ton/ ha (Alemayehu Mengistu, 2006a; Alemayehu Mengistu and Getnet Assefa, 2012)

In Ethiopia, most improved tropical species can grow in the areas from 1500–2000 meters above sea level and temperate species grow from above 2100 up to 3000 meters above sea level (Alemayehu Mengistu, 2003). Introduced improved forage yield is higher than the naturally occurring swards and has higher nutritional value. Besides, the length of the productive season is longer for cultivated pastures than for native pastures, which provides

9

an opportunity for dairy and fattening production to develop and use pasture and forage on a large scale. Greater use of leguminous fodder trees and shrubs assists in increasing soil fertility, controlling soil erosion and providing firewood and timber. Fodder crops such as oats and vetch mixtures, alfalfa, Rhodes grass, and fodder beet are commonly grown at higher altitudes for livestock feeding although the area coverage is minimal. Fodder crops have had minimal use in non-dairy production, perhaps partly because a seed has been imported and available only in limited quantities. In suitable areas, yields of oats/Vetch mixtures are commonly 8-12 tons of dry matter per hectare (Alemayehu Mengistu, 2006a). In most parts of country, introduced fodder trees (Leucaena, Sesbania sp. tree Lucerne) and grasses like Napier grass, Oat ,Rhodes grass, desho grass, etc., legumes like Dismodium, Vetch, sweet lupin are on the way of production, These forages had been used mostly within the soil-erosion control program, irrigation canals, around farmers' homesteads and in small amount in the crop cultivated lands are some of forage production strategies in Ethiopia. The utilization strategies of these improved forages are used as cut and cry system mostly, controlled grazing, sometimes as fresh feed and making hay. Nutritive value of some of improved forages species in Ethiopia are listed below (Table 2.2).

Table 2. 2 Avarage nutritive value some improved forages available in Ethiopia (%)

Feed DM OM CP NDF ADF ADL IVDM IVO ME Source (%) (%) (%) (%) (%) (%) D (%) MD (Mcal (%) /kg) Alfalfa hay 90.58 87.23 23.7 25.55 20.5 3.55 72.45 - 2.45 Ethiopian Feed Database Cow pea hay 90.9 84.45 21.5 37.85 34.4 6.37 73.31 - 2.37 available online Lablab hay 91.08 91.23 17.1 47.3 38.8 8.32 66.01 - 2.35 (http://www.vslp Napier grass hay 92.04 82.87 7.47 64.28 41.1 5.51 64.98 - 1.22 .org/ETHFeed/de fault.asp) Pigeon pea hay 94.44 - 16.2 54.64 40.5 14.4 53.46 - 1.91 Rhodes grass hay 91.48 88.01 7.33 70.24 32.4 5.04 60.61 - 2.05

Desho grass hay 30.16 89.78 7.86 71.81 42.5 5.37 - 43.9 1.477 Bimrew Asmare et al. (2017) Vetch hay 91.08 89.5 18.1 36.17 46.6 8.36 69.25 - - Kassahun Desalegn and Wasihun Hassen,2015 Tree Lucerne hay 91.79 95.09 22.6 59.65 37.9 68.1 Shigdaf Mekuriaw et al., 2018 10

2.1.4. Agro-industrial by-products

Agro-industrial by-products have special value in feeding livestock mainly in urban and peri urban livestock production system, as well as in situations where the productive potential of the animals is relatively high and require high nutrient supply. The major feed resources in the country are natural pasture and crop residues, with agro industrial by-products and manufactured feed contributing much less (Berhanu Gebremedihn et al., 2009). Natural pasture and crop residues are naturally of low quality and do not fulfill the nutrient requirement of animals. Hence, high producing animals such as dairy cattle and fattening animals should be supplemented with high energy and/or protein concentrates. The most commonly used concentrate feeds in Ethiopia belong to different agro-industrial byproducts including: Milling byproducts such as wheat bran, Wheat middling, oilseed cakes such as noug cake, cottonseed cake, peanut cake, linseed cake, sesame cake, sunflower cake molasses and bagasse from sugar factories, brewery byproducts and occasional surplus grain or grain damaged during processing (Fekede Feyissa et al., 2011; CSA, 2018). Further, agro- industrial by-products produced in Ethiopia include by-products from flour milling, sugar factory, and oil processing factories, abattoir, and breweries. These products are mainly used for dairy, fattening and commercial poultry production and the scope for their wider use by smallholder producers is low due to availability and price (Yoseph Mekasha et al., 2010).

Table 2. 3 Average nutritive value of agro-industrial by-products available in Ethiopia

Feed DM OM ADF NDF ADL CP IVDM ME ( P Ca (%) (%) (%) (%) (%) (%) D (%) Mcal/kg) (%) ( % ) Wheat bran 89.3 95.0 13.6 46.4 3.3 16.8 - - 1.07 0.14 Wheat middling 89.7 94.9 23.5 42.8 4.5 16.3 61.3 2.2 0.94 0.14 Noug cake 92.3 89.7 31.5 37.6 12.4 31.4 61.3 2.1 1.15 0.76 Cotton seedcake 92.0 92.2 20.7 38.4 6.3 41.0 70.4 2.5 1.52 0.30 Groundnut cake 92.7 95.0 8.7 16.4 1.7 55.5 80.9 3.0 0.61 0.14 Linseed cake 91.5 93.3 19.2 31.2 7.3 29.6 70.3 2.6 0.84 0.51 Sunflower cake 91.8 94.0 31.0 35.5 9.7 25.9 58.9 2.2 1.02 0.63 Sesame cake 91.4 88.6 27.8 32.7 4.9 25.5 79.3 2.6 0.77 1.31 Rape seedcake 90.9 91.1 20.0 25.8 8.1 35.8 72.3 2.6 1.19 0.73

(Source: Seyoum Bediye, et al., 2007, cited by Firew Tegegne and Getnet Assefa, 2010; Ethiopian Feed Database available online (http://www.vslp.org/ETHFeed/default.asp) 11

2.1.5. Other feed resources

Livestock feed sources are classified as conventional and non-conventional, where the non- conventional ones vary according to feed habit of the community and others, e.g. vegetable refusals are non-conventional (Alemayehu Mengistu, 2003). Non-conventional feed sources are generally refer to all those feeds that have not been traditionally used for feeding livestock and are not commercially used in the production of livestock feeds. Non- conventional feeds such as vegetable refusals, sugarcane leaves, Enset leaves, fish offal, left over of Enjera and Porridge, home waste, local brewery by products etc. are used as animal feed in the different parts of Ethiopia (EndaleYadessa et al., 2016). Thus non-conventional feeds could partly fill the gap in the feed supply, decrease competition for food between humans and animals, reduce feed cost, and contribute to self-sufficiency in nutrients from locally available feed sources. However, non-conventional feeds are not available at large and their contribution to livestock feed as copping strategy was very small proportion (1.61%) (CSA, 2018).

2.2. Overview of Brachiaria grass

2.2.1. Historical background of Brachiaria grass

The genus Brachiaria and its species are originated from Africa but found their way to the subtropical and tropical of Australia and South America (Pizarro et al., 2013). In Brazil, the introduction of Brachiaria grass species stems 500 years back when Brachiaria Mutica was introduced through African slaves (Miles et al., 2004). The new forage grasses were favored by stock owners due to their persistence under grazing and higher nutritional value compared to indigenous grasses. Brachiaria species are common constituents of the natural vegetation in East Africa. Within these areas, cut and carry or extensive grazed pasture is often practiced. Miles et al. (2004) state that the forage potential of these grasses was first recognized about 40 years ago in restricted niches in tropical Australia. The major impact of the genus was, however, realized between 1976-1981 when a handful of germplasm was sown in Tropical America and numerous cultivars have been developed to date.

Brachiaria, native African grass and it welcome back home, after its improvement in Brazil (South America) and Australia. It is an important constituent of natural vegetation across 12

sub-Saharan African. Brachiaria grass consists of over 100 species, but only seven species are widely evaluated as cultivated pasture in tropical Africa (Vendramini et al., 2011). These are perennial species of African origin B. arrecta, B. brizantha, B. decumbens, B. dictyoneura, B. humidicola, B. mutica and B.ruziziensis have been used as fodder plants, particularly in tropical America It is the most extensively cultivated forage in South America, Australia, and East Asia. However, the use of Brachiaria as cultivated forages in Africa is extremely limited. Brachiaria grasses are among the most important tropical grasses that originated from Africa, improved in Americans through agronomic selection and breeding (Miles et al., 2004) and demonstrated to be highly productive, nutritive and socially acceptable in Asia and Africa for different livestock production systems (Mutimura and Everson, 2012; Pizarro et al., 2013; Vendramini et al., 2014).

2.2.2. The Biology and Agronomy of Brachiaria

Brachiaria are C4 grass some of which are annual and others are perennial. They belong to the Poaceae family. Brachiaria species are adapted to a wide range of soil types, from Oxisols and Ultisols (low-fertility acid soils) to Alfisols and Mollisols (high-fertility neutral soils) and perform much better on acid soils than other grasses such as panicum species (Ghimire et al., 2015). Brachiaria is drought tolerant and adapted to low fertility soils of sub-Saharan Africa; it can play a significant role in soil fertility improvement, soil conservation, increasing biodiversity and minimizing greenhouse gas emissions (Rao et al., 2014). Brachiaria grows well in different agro-ecological zones of tropical Africa. It performs best in sub-humid and humid environments where rain falls surpass 700mm and mean temperature exceeds 19 0c. Lower temperature slows down the growth rates, hence the Brachiaria grass performs poorly 1800m above the sea level. It grows on a wide range of soil types including those of low fertility. Locations that longer dry season of over five months is not suitable for Brachiaria unless there is provision for irrigation (Vendramini et al., 2011). The biomass yield and nutritional attributes of Brachiaria grass depend on plant management. In tropical areas concerning phenology, soil fertility, moisture conditions, light intensity and temperature (Cook et al.,2005; Campos et al., 2013), the most sensitive attributes to management and environment are metabolizable energy and crude protein contents as well as macro and micro-minerals. Brachiaria grass produces a lot of nutritious biomass as much as 15 tons of dry matter per acre/year, which in turn, when fed livestock increases milk and meat production (Vendramini et al., 2011). 13

2.3. Brachiaria grass in Ethiopia

Brachiaria grass species are one of the most important tropical grasses distributed throughout the tropics especially in Africa and has been introduced in some areas of Ethiopia as animal feed, part of push-pull technology and for soil conservation by different organizations, i.e., ICIPE, LIVES Project, Wollo University and MoANR (Wubetie Adnew et al., 2018a). In Ethiopia the production and productivity of Brachiaria, the grass is at the infant stage, since its introduction in the country five years ago (ICIPE, 2017). Recently Wubetie Adnew et al.(2018a) survey result in selected areas of Amhara and Oromiya indicated that Brachiaria grass information and utilization was at infant stage which needs detail researches from adaptation on small plot of land to large scale, including evaluation on animal performance though of different animal class at different management practices and environmental conditions. Wubetie Adnew et al. (2018b) reported on wild Brachiaria brizantha grass ecotypes in northwestern Ethiopia, the overall results obtained in the three agro-ecologies (lowland, midland and high land) an adaptive and productive at different harvesting at 60 to 90 to 120 days, even if, the variations were observed between ecotypes. On the other hand Mimila Zemene (2018) also reported on Brachiaria mutica grass in the northwestern Ethiopia. This Brachiaria grass was adaptive and productive under different harvesting and spacing (60 days to 90 days to 120 days and 15, 30 and 45 cm, respectively as reported by the same author. Similarly, Negasu Yibarkew (2018) reported on Brachiaria hybrid Mulato II in low land northwestern Ethiopia. As indicated the production and productivity in terms of forage yield and chemical composition of this Mulato II was affected by type of fertilizer and plant spacing.

In Africa the largest uptake of hybrid Brachiaria cv. Mulato-II is currently taking place in eastern Africa, where the grass is used as a trap plant in the push-pull system that helps control maize stem borers and the parasitic weed, Striga hermonthica (Khan et al., 2014). The push-pull-system had developed and promoted by the International Centre of Insect Physiology and Ecology (icipe) (Khan et al., 2014). This smart technology successfully harnesses agro-biodiversity for improving the productivity of cereal crops while providing fodder for livestock. Initially, its components included Napier grass and Silver leaf desmodium (Desmodium uncinatum). Over 20,000 smallholder farmers benefiting from the adopt project in Kenya, Uganda, Tanzania, Nigeria, and Ethiopia have already planted cv. Mulato-II. Farmers in Kenya indicated that their dairy milk production has doubled due to 14

the availability of the improved Brachiaria grass. They prefer cv. Mulato-II to Napier grass for several reasons: it is drought-tolerant, highly palatable and nutritious for livestock, easier to handle as cut-and-carry and for making hay to use during the dry season. As the push- pull-system has developed to control maize stem borer, little attention has paid thus far to the possible importance of livestock production improvements for the uptake and further spread of the technology (Maass et al., 2015).

2.4. Descriptions of specific Brachiaria grass cultivars

2.4.1. Mulato-II hybrid

Mulato II is the result of three generations of crosses and screening conducted by the International Center for Tropical Agriculture (CIAT) in Colombia, including original crosses between Brachiaria ruziziensis R. Germ. &Evrard clone 44-6 (sexual tetraploid) x Brachiaria decumbens Stapf cv. Basilisk (apomictic tetraploid). Sexual progenies of this first cross were exposed to open pollination to generate a second generation of hybrids. From the second generation of hybrids, a sexual genotype was selected for its superior agronomic characteristics and again crossed to produce (Argel et al., 2007). Mulato-II is a semi-erect perennial apomictic grass that can grow up to 9 feet tall. Seed establishes it, although it could be propagated vegetatively with stem segments. It produces several cylindrical stems (some with a semi-prostrate habit) capable of rooting at the nodes when they come into contact with the soil. They have lanceolate and highly pubescent leaves of 35 to 65 cm in length and 2.5 to 4 cm width (Guiot and Melendez, 2003). This grass species are also adapted to many soil types, ranging from sands to clays; however, it does not tolerate poorly drained soils (Vendramini et al., 2010). It offers considerable potential to alleviate feed shortage and complement the feeding value of other grasses by introducing in the farming systems. Its benefits as a livestock feed have been quantified clearly in humid and sub-humid climate (Urio et al., 2006, Hare et al., 2007). Mulato-II is a vigorous, semi- erect grass species with very deep and branched roots giving it superior drought resistance and it holds the key to improve livestock productivity in the tropics (Pizarro et al., 2008)

2.4.2. Brachiaria mutica

B.mutica (Para grass) is a common weed in many cane-growing areas. It has been used in tropical locations as a fodder species, especially as a pounded pasture in beef production. 15

This plant can be a very aggressive invader, particularly in low-lying ungreased areas and in sugar cane crops. It is often found in wet situations, especially drains, but will also grow in deep soils in non-swampy areas. Brachiaria mutica, known as para grass (Africa, Australia) or buffalo grass is under the tribe Paniceae and subfamily Panicoideae of family Poaceae. Poaceae, also known as gramineae is one of the largest families in the world. In India, this family has 263 genera and 1291 species. It is the largest family of India. Among these genera, B. mutica is very common in West Bengal as well as in India. B. mutica is sometimes known as Urochloa mutica (Forsskil) T.Q. Nguyen. This plant is an important plant of tropical used as green foliage, hay or browse (Saurav and Amal, 2011). They are also useful for banks of streams to control erosion. B. mutica (Forsskil) Stapf is a creeper, decumbent, stoloniferous, perennial grass with long coarse stolons. Culms are semi- erect, greenish, cylindrical, node and internodes prominent, internodes solid, glabrous, node swollen and soft. Culm at first spread horizontally then upward. Sheath is mostly longer than internodes, papillose, pubescent, highly hairy. Hairs are unicellular and cylindrical in type (Saurav and Amal, 2011).

2.4.3. Brachiaria brizantha

Brachiaria brizantha (Palisade grass) is a highly productive grass that is propagated by seed and vegetatively by clumps and stems. It requires more fertile and better-drained soils than other species of Brachiaria and has a higher tolerance to drought. Palisade grass persists under severe grazing and frequent harvesting (Vilela et al., 2004). It can spread and suppress weeds and is highly resistant to rust, leaf-cutting ants (Atta cephalotes) and spittlebugs (Deois sp and Notozulia entreriana), but is highly susceptible to Rhizoctonia foliar blight (CIAT, 2001). Palisade grass is one of the most cultivated forage grasses in Central Brazil, due mainly to spittlebug resistance and high yield potential. In the Zona da Mata of the state of Pernambuco, northeastern Brazil, the total herbage accumulation in palisade grass pastures can reach 28 Mg dry matter (DM) ha-1 during the grazing season ( Santos et al., 2003). Marandu is palisade grass currently ranks first in the Brazilian forage seed market: 44% of the total amount of seed commercialized (Vilela et al., 2004). It is tufted, prostrate or semi-erect perennial with short rhizomes and stems 30-200 cm tall.Leaf linear to broadly linear, 10-100mm, glabrous or hairy. Two cultivars have been released in South America: the best known and most widespread is "Marandú" (Brazil) where as "La Libertad" is a recent Colombian release. Like signal grass, palisade grass is well adapted to the humid and 16

sub-humid tropics where it can withstand dry seasons of up to 5 months. It grows well on a range of soils including sandy and acid soils, but it requires more fertile soil than signal grass. It does not tolerate poorly drained soils. Like signal grass, it tolerates light to moderate shade (Vilela et al., 2004).

2.4.4. Mulato-I hybrid

Mulato-I was resulted from crossing Brachiaria ruziziensis clone 44-6 X Brachiaria brizantha CIAT 6297 and this was carried out in 1988 (CIAT,2001). Mulato is a semi-erect perennial apomictic grass that can grow up to 1.0 m tall. It established by seed and although rooted stem stocks could propagate it vegetatively. It produces vigorous cylindrical stems, some with a semi-prostrate habit, capable of rooting at the nodes when they come in close contact with soil. Mulato-I has lanceolate and highly pubescent leaves of 40 to 60 cm in length and 2.5 to 3.5 cm in width (Guiot and Melendez, 2003). Mulato-I grows well in humid tropical areas with high rainfall and short dry periods, and in sub-humid conditions with 5 to 6 dry months and an annual rainfall of 700 mm. (Argel et al., 2005) reported it that Mulato-I grows well in subtropical conditions where periodic frost occurs, such as southern Florida in the USA. It grows in acid to alkaline soils (pH 4.2-8.0), but it requires medium to high fertility and good drainage. Mulato-I is drought tolerant and can regrow again during critical times of the year. It has CP concentration fluctuating between 90 to 170 g kg-1 and digestibility of 550 to 620 g kg-1 (CIAT, 2005; CIAT, 2006).

2.4.5. Brachiaria decumbens

Brachiaria decumbens is a low-growing decumbent perennial grass with flowering stems up to 100 cm high originating from the prostrate, multi-nodded stems; plants can spread by both rhizomes and stolon’s as well as through seed production. Brachiaria decumbens has been widely adopted as a pasture for grazing ruminants due to its high nutritive value and aggressive growth habit that provides a dense ground cover able to suppress weeds. It is resilient and is grown over a wide range of soil types including infertile acid soils with low pH (<3.5) and high aluminum levels and climates ranging from tropical to sub-tropical. In Western Australia, Brachiaria decumbens was included in a mixed-species pasture mix for the sub-tropics (sandy soils and a maximum annual rainfall of around 600 mm (Moore, G et al, 2013). B.becumbens produces more dry matter than most tropical grasses during the

17

dry season and is capable of producing 15–27 MT (metric tons) dry matter (DM)/hectare/year (Abdullah, A.S, 2015). In a study undertaken in South Sulawesi, Indonesia, Brachiaria decumbens produced more dry matter than other improved pasture species during the dry season (5700 kg/ha and 1910–3500 kg/ha respectively). It has been suggested that the ability to respond to small amounts of rainfall that occurred in the dry season was due to the extensive root system of Brachiaria decumbens, plants produce new growth rapidly with out-of-season rain events during the dry season and with the break of the season. Young vegetative material has been associated with reported outbreaks of photosensitization among ruminants grazing Brachiaria decumbens pastures (Riet-Correa, 2007). Brachiaria decumbens has shown, in plot trials, an increasing response to the addition of nitrogen fertilizer up to 400 kg N/hectare depending on nitrogen source( Teixeira, et a., 2011), producing more than 40 MT DM/hectare/year at this level. In Brazil, the late summer application of nitrogen (100 kg N/ha) together with 95 days of deferred grazing, was found to produce the greatest biomass on an annual basis. Split applications of nitrogen fertilizer have been recommended as the growth response is greater in the period immediately following the application of nitrogen (Teixeira et al., 2011).

2.5. Effect of fertilizer, cultivars and soil type on plant morphological characteristics

2.5.1. Plant height

Plant height is the distance between the upper boundary of the main photosynthetic tissues (excluding inflorescences) on a plant and the ground level. Growth parameters play a vital role in enhancing fodder yield (Imran et al., 2007). Plant height as an important yield parameter and it suggested that canopy of plant height could be considered a valuable tool to make a rough estimation of forage dry matter yield. Plant height is an important parameter contributing to biomass yield in forage crop and it is an important component, which helps to determine the growth attained during the growing period (Tessema Zewdu et al., 2003).

Previous reports indicated that the height of the plant was dependent on the level and type of fertilizer. According to report of Worku et al. (2017) in Jinka Agricultural Research Center, Southern Ethiopia indicated that plant height of desho was increased as increase of fertilizer level from 0.9 m to 1.3 m at 0.5 kg and 0.5kg/ha DAP and UREA to 150 & 200

18

kg/ha respectively. Similarly Biniyame Mihret et al. (2018) at Injebera district in northwestern Ethiopia, reported that mean plant height of desho grass significantly affected by fertilizer type and level of fertilizer, of which the height plant was recorded from NPS fertilizer (55.23 cm) than manure (52.22 cm) and without fertilizer application (42.01 cm). Nafatul et al. (2015) reported that as the level of NKP fertilizer increased (0,150 and 300 kg/ha), the mean plant height of Brachiaria brizantha Marandu was also increased from 6.74, 7.86 and 10.13 cm respectively at 60 days of harvesting age which was established by seed propagation system. This shows the application of fertilizer on the different levels and types could be a critical factor for affecting plant height.

On the other hand, the type of species and cultivars could affect plant height. Wubetie Adnew et al. (2018b) reported on Brachiaria brizantha grass ecotypes (Eth. 13726, Eth. 13809 and Eth. 1377) in northwestern Ethiopia. The overall mean of the tallest plant height was recorded from Eth. 1377 (58.73 cm) followed by Eth. 13809 (51.47 cm) and Eth. 13726 (51.07 cm). This indicates that plant height can be different among different accessions of Brachiaria species or cultivars with the same management system. The overall results obtained in the three agro-ecologies, lowland (60.5 cm), midland (57.62 cm) and high land (43.15 cm) indicated that plant height was significantly affected by agro ecology and harvesting age which is increasing trends from 60 to 90 to 120 days, even if, the variations were observed between ecotypes. Similarly, Susan et al. (2016) on the primary production variables of Brachiaria grass cultivars in Kenya dry lands reported that at week 16, Napier recorded the highest mean plant heights (103.8cm) and at Llanero lowest (6cm). Among the Brachiaria cultivars, MG4 (63.4cm) recorded higher plant heights and although second after Napier (103.8cm), its height was not significantly different from C.gayana cv. Kat R3 (52.8cm). Mustaring et al. (2014) indicated that Brachiaria mutica had the highest plant height (207.47cm) than Brachiaria brizantha and Brachiaria Mulato-I at 8 weeks of harvesting. This shows that Brachiaria has tremendous potential to increase the availability of forage dry matter in the livestock feeding system to overcome shortage of feed.

Regarding soil type effect on plant height of forage grass was limited information on Brachiaria grass cultivars or related grass families in different regions including Ethiopia. However, Getnet Assefa and Inger Ledin (2001) reported that the overall plant height of Oat cultivar was recorded at red soil (155 cm) significantly higher than black soil (110 cm). This shows that soil type had the effect on the plant height. 19

2.5.2. Number of tiller per plant

Tillers are new grass shoots, made up of successive segments called phytomers, which are composed of a growing point (apical meristem which may turn into a seed head), a stem, leaves, roots nodes, and latent buds; all of which can arise from crown tissue buds, rhizomes, stolon's, or above ground nodes (aerial tillers). Tillers are very important in understanding grass growth and regrowth.

Previous reports were indicated that number of tillers per plant (NTPP) were affected by the level and type of fertilizer. Worku et al. (2017) in Jinka Agricultural Research Center, Southern Ethiopia indicated that there was a significant difference (p<0.05) in the number of tillers per plant of desho grass as the fertilizer rate increased from 100 kg/ha to 250 kg/ha which was 43.6 and 79.6 tillers per plant respectively. On the other hand, Kizima et al. (2014) reported that the application of the optimal level of Nitrogen fertilizer significantly affects the appearance of new tillers and increases the dynamics of the tiller population of Cenchrus ciliaris. Biniyame Mihret et al. (2018) in northwestern Ethiopia reported that number of tillers per plant of desho grass was significantly affected by fertilizer type and level of fertilizer, of which the number of tillers per plant was recorded from NPS fertilizer (61) higher than manure (50) and without fertilizer application (37). This indicates that the application of different levels and type fertilizers are a critical factor in the number of tillers per plant.

On the other hand, the type of species affected the number of tillers per plant /variety and cultivars as indicated below by different authors at different locations. Wubetie Adnew et al. (2018b) reported that Brachiaria brizantha grass ecotypes of Eth. 13726, Eth. 13809 and Eth. 1377 in northwestern Ethiopia. The overall mean of the maximum number of tillers per plant was counted from Eth. 13809 (33.53) followed by Eth. 1377 (29.38) and Eth. 13726 (26.35). The overall number of tillers per plant was obtained in the three agro-ecologies were 33.62, 28.52 and 27.12 for midland, lowland, and highland respectively. NTPP were significantly affected by type of ecotypes and harvesting age which is increasing trends from 60 (17.89) to 90 (25.07) to 120 (36.09) days tiller per plant, even if, the variations were observed between ecotypes. Mustaring et al. (2014) reported that tiller number per plant increased as maturity, which Brachiaria Mulato-II (122.4 per plant) had the highest (P<0.05) tiller number than Brachiaria brizantha (63.73) and Brachiaria mutica (57.7) at

20

similar harvesting age (60 days) in lands of central Sulawesi, Indonesia. Overall indicates that the number of tillers per plant are affected by the type of species or cultivars at the same management practices and similar environmental conditions.

Regarding soil type effect on number tillers per plant of forage grass was limited information on Brachiaria grass cultivars or related grass families in different regions including Ethiopia. However, Getnet Assefa and Inger Ledin (2001) reported that the overall tillering number of Oat cultivar was recorded at red soil (454) significantly higher than black soil (365) at a one-meter square area (1m2). This shows that soil type have effect the number of tillers per plant. Mimila Zemene (2018) on Brachiaria mutica grass at Bahir Dar, Ethiopia reported that the maximum number of tiller per plant (67.4) was obtained from late harvesting age (120 days) followed by 60 and 90 days of harvesting, which produce 57.8 and 57.9 tiller numbers respectively at wider spacing, which was 45 cm between rows.

2.5.3. Number and length of leaves per plant

Leaf length per plant (LLPP) is a key factor determining the vegetative yield of forage grasses. Leaf length in grasses plays an essential role in shaping the physical structure of the canopy and consequently on the competition for the light within the plant. One of the major adaptive responses to light competition in plants is an increase in plant height, i.e., leaf length during the vegetative period in grasses (Barre et al., 2009).

The number leaves per plant (NLPP) and length of leaves per plant (LLPP) were affected by the type and level of fertilizer. Biniyame Mihret et al. (2018) in northwestern Ethiopia, reported that the number of leaves per plant and length of leaves per plant of desho grass were significantly affected by fertilizer type. The mean of leaf length was 30.48 cm,27.37 cm and 19.87 cm for NPS, manure and without fertilizer respectively, while the mean number of leaves per plant were highly decreased from NPS (530) to manure (388) to without fertilizer (307). Nafatul et al. (2015) reported that as the level of NKP fertilizer increased (0 to 150 to 300kg/ha) the mean leaf length per plant of Brachiaria brizantha Marandu was linearly increased from 12.32 to 12.4 to 14.02 cm respectively at 60 days. While the number of leaves per plant were increased from 4.87 to 6.62 to 10.12 leaves per plant at 0,150 and 300 kg/ha NKP fertilizer levels respectively, which was established by seed propagation system. This application of fertilizer in the different levels and types are a

21

critical effect on the number and length of leaves per plant at the same management practices as well as similar environmental conditions.

On the other hand, the number and length of leaves per plant were affected by the type of acesstions. Wubetie Adnew et al. (2018b) reported on Brachiaria brizantha grass ecotypes of Eth. 13726, Eth. 13809 and Eth. 1377 in northwestern Ethiopia. The overall mean of a maximum number of leaves per plant was recorded from Eth. 1377 (5.01) followed by Eth. 13726 (4.74) and Eth. 13809 (4.36). While the longest length of leaves per were measured from Eth. 13809 (20.1 cm) followed by Eth. 13726 (17.29 cm) and Eth. 1377 (16.22 cm). The overall number of leaves per plant were obtained in the three agro-ecologies is 5.04, 4.96 and 4.11 for midland, lowland, and highland respectively. While the overall mean longest leaf length per plant was recorded from midland (20.22 cm) followed by lowland (18.93 cm) and highland (14.47 cm). This implies that the length and numbers of leaves per plant are different among accesstion genetic variation at similar management practices. Regarding soil type effect on the number of leaves per plant and length of leaf per plant of forage grass were limited information on Brachiaria grass cultivars or related grass families in Ethiopia.

2.5.4. Leaf to stem ratio

Leaf: stem ratio (LSR) of grass is an important factor affecting diet selection, intake by animals and overall nutritive value forage. Estimates of leaf: stem ratios were commonly based on a labor-intensive process of hand separating leaf and stem fractions (Smart et al., 2004). The reduced leaf-to-stem ratio is a major cause of the decline in forage nutritive value with maturity, and the loss in nutritive value that occurs under adverse hay curing conditions. Leaf to stem ratio reflected the variation of leaf stem mass with harvest and is a trait that can affect preference during grazing. Leaves are higher in nutritive value than stems and the proportion of leaves in forage declines as the plant matures (Ball et al., 2001).

There is limited information on the leaf to stem ratio in the different species of grasses or cultivars as affected by level and type of fertilizer, however, some information was available on the leaf to stem ratio as affected by harvesting age and plant spacing. Mimila Zemene (2018) reported on Brachiaria Mutica grass indicated that harvesting date and spacing affect the productivity of grass. Narrow plant spacing (15cm) at early harvesting age (60days)

22

resulted in significantly higher (1.37) leaf to stem ratio compared to intermediate (90days) and late harvesting (120days) and it was statistically similar with intermediate (30days) and wider plant spacing (45cm) at early harvesting (60days). Similarly, Bimrew Asmare et al. (2017) reported that early harvesting (90 and 120 days) resulted in significantly higher leaf- to-stem ratios compared to the late harvesting date (150 days). Chloris gayana KATR3, on the other hand, has a higher proportion of stem relative to leaf by week sixteen, which could be the reason for lower dry matter yields (Nguku et al., 2016). Leaf to stem ratio values of Panicum coloratum were significantly affected by the age of regrowth (Driba Geleti and Adugna Tolera, 2013). Regarding soil type effect on the leaf to stem ratio of forage grass was limited information on Brachiaria grass cultivars or related grass families in different regions including Ethiopia. However, Getnet Assefa and Inger Ledin (2001) reported that the leaf to stem proportion of oat cultivars was recorded at black soil (40.9 leaf to 37.3 stem) higher than red soil (37 leaf to 39.7stem).

2.5.5. Number and length of roots per plant

Roots provide essential functions including the uptake of water and nutrients for plant growth, serve a role as storage organs, anchor the plants to the soil and are the site of interactions with pathogenic and beneficial organisms in the rhizosphere (Paez et al., 2015). An increase in root length increase the absorbing surface, thus may be important for water and mineral absorption. The explanation was that the uptake of a different nutrient in competing plants was proportional to the root length, as the greater the root length, the shorter the distance the nutrient has to travel to the root (Crush et al., 2010). Deeper roots provide plants with better access to stored water and nutrients such as N, a soluble nutrient that tends to leach into the deeper layers of the soil (Wasson et al., 2012). Root mass is heritable in perennial ryegrass at levels that will allow breeding for larger root systems (Crush et al., 2006), and progress is now being made towards changing the shape of root systems towards deeper rooting patterns.

The number of roots per plant (NRPP) and length roots per plant (LRPP) were affected by the level and type of fertilizer. Biniyame Mihret et al. (2018) reported that the fertilizer type significantly affects the number of roots and length roots of desho grass in highlands of northwestern Ethiopia. The highest mean number of roots per plant were recorded for NPS (116) followed by manure (94) and while the lowest was from without fertilizer (71). The

23

same author reported that the length of roots per plant were significantly affected by fertilizer type which is the highest mean was observed from NPS (26.05 cm), manure (25.36 cm) and the lowest was observed from without fertilizer or control group (22. 84 cm) at 120 days of harvesting stage. Regarding soil type effect on number and length of roots per plant of forage grass was limited information on Brachiaria grass cultivars or related grass families in different regions including, Ethiopia. The number of roots and length of roots per plant were affected by the harvesting stage and plant spacing. Mimila Zemene (2018) reported on Brachiaria Mutica grass indicated that the highest number of roots per plant (208.5) were counted from wider plant spacing (45cm) and the late harvesting age (120days) followed by (139.7) number of roots per plant at 90 days of harvesting age. While the lowest number of roots per plant (99.3) were at 60 days of harvesting age. Therefore, the number of roots increases linearly with the maturity of the plant.

24

Table 2. 4 Summary on morphological characteristics of Brachiaria grass species by different authors

Species HD ALT PH(cm) NTPP LSR NLPP LLPP Source B.brizantha 8 Mid 145.43 57.57 - - Mustaring et al.,2014 B.butica 8 Mid 207.47 63.73 - - Mustaring et al.,2014 B.mutica 8 Mid 105 40 1.27 249 19 Mimila, 2018 B.mutica 12 Mid 180 48 1.03 616 22.6 Mimila, 2018 B.mutica 16 Mid 255 565 0.97 1038 26.2 Mimila, 2018 B.Mulato-I 8 Mid 117.23 122.4 - - Mustaring et al.,2014 B.Mulato-I 8 Mid 14.2 52 - - - Susan et al., 2015 B.Mulato-I 12 Mid 28.2 51 - - - Susan et al., 2015 B.Marandu 8 Mid 19.2 44 - - - Susan et al., 2015 B.Marandu 12 Mid 23.1 36 - - - Susan et al., 2015 B.La Liberated 8 Mid 24 46 - - - Susan et al., 2015 B.La Liberated 12 Mid 75.6 74 - - - Susan et al., 2015 B.decumbens 8 Mid 26.7 74 - - - Susan et al., 2015 B.decumbens 12 Mid 80.1 56 - - - Susan et al., 2015 Mulato-II 12 Low 65 30 - - - Clara ,2013 Mulato-II - Low 73.93 26.76 1.15 251.87 31.72 NegasuYibarkew, 2018 Eth. 13726/ B.brizantha ecotypes 12 Low 59.39 24.71 - 4.64 18.61 Wubetie et al., 2018b Eth.13809/ B. brizantha ecotypes 12 Mid 56.97 36.67 - 5.51 19.39 Wubetie et al., 2018b Eth. 1377/ B.brizantha ecotypes 12 High 38.86 20.67 - 4.08 13.39 Wubetie et al., 2018b

HD= Harvesting date, ALT= Altitude, PH= Plant height, NTPP= Number of tillers per plant, LSR= Leaf to stem ratio, NLPP=Number of leaves per plant, LLP= Length of leaves per plant

25

2.6. Effect of fertilizer, cultivars and soil type on forage dry matter yield

Dry matter yield (DMY) per unit of area is the most important parameter for livestock feed. It can calculate by multiplying dry matter percentage, green forage yield, and area of production. Forage dry matter yield was affected by level and type of fertilizer. Biniyame Mihret et al. (2018) in highlands of northwestern Ethiopia reported that DMY of desho grass was significantly affected by fertilizer type. The highest mean DMY ton per hectare at NPS (24.08) followed by manure (14.05), while the lowest DMY was at control or without fertilizer (8.55). Worku et al. (2017) indicated that there was significant difference (p<0.05) in DMY of desho grass ton per hectare was increased as the fertilizer rate increases from 100 kg/ha (8.32 ton /ha) to 150 kg/ha (9.09 ton/ha) to 200ka/ha (12.73 ton /ha to 250 kg/ha 13.6). Similarly Nafatulk et al. (2015) reported that as the level of NKP fertilizer increased from (0,150 and 300 kg/ha) the mean DMY of Brachiaria brizantha Marandu was linearly increased from 0.78, 1.13 and 0.99 ton per hectare respectively at 60 day harvesting age fertilizer levels, which was established by seed propagation system. This indicates that the application of fertilizer with the different levels and types are critical factors for forage dry matter yield under normal and optimum recommendations.

On the other hand, it could also showed variations in DMY among different species or cultivars of grasses family. Wubetie Adnew et al. (2018b) on Brachiaria brizantha grass ecotypes in northwestern Ethiopia reported that the overall highest DMY was recorded from Eth.1389 (5.91 t/ha) followed by Eth.1377 (5.01 t/ha) and while lowest was from Eth.13726 (4.05 t/ha). The overall highest DMY was recorded from lowland agro ecology (5.93 t/ha) followed by midland (5.1) while the lowest DMY was at highland agro ecology (4.02 t/ha). Similarly, Susan et al. (2015) noted that DMY kg per hectare was significantly affected by the type of cultivars at different harvesting ages. Similar reports by Mustering et al. (2014) what types of species and cultivars significantly affected DMY kg per hectare. Regarding the effect of soil type on the dry matter yield of Brachiaria grass is limited. However, Getnet Assefa and Inger Ledin (2001) reported that the dry matter yield of Oat is significantly higher at red soil (9.39 t/ha) than black soil (5.45 t/ha), even if the variations were observed among Oat cultivars. This indicates that soil type is a critical factor for forage productivity in terms of forage dry matter yield and nutritive value.

26

Table 2. 5 Summary on dry matter yield of Brachiaria grass species by different authors

Brachiaria Species Parameters ( ton /ha) Harvesting age (week) Agro ecology Source 8 12 16 Low Mid High B.Brizantha 5.1 - - - 5.1 - Mustaring et al.,2014 B.Mutica 5.5 - - - 5.5 - Mustaring et al.,2014 B.Mutica 5.5 13.43 18.4 - - - Mimila, 2018 B.Mulato-I 7.9 - - - 7.9 - Mustaring et al.,2014 B.Mulato-I 4.9 8.6 - - - Susan et al.,2015 B.Mulato-I - 5.0 - Mutimura and Everson (2012) B.Mulato-II - 2.94 17.3 Low - - Clara ,2013 B.Mulato-II - 5.1 Mutimura and Everson (2012) B.Marandu 5.5 5.36 - - Susan et al., 2015 B.Mulato-II - 14.59 - Negasu Yibarkew ,2018 B.La Liberated 5.85 8.015 - - - - Susan et al.,2015 B.decumbens 5.95 6.7 - - - Susan et al., 2015 B.decumbens - 4.8 - Mutimura and Everson (2012) Eth.13726/B.brizantha ecotype 1.69 3.63 6.83 4.24 4.6 3.28 Wubetie et al., 2018b Eth.13809/B.brizantha ecotype 2.95 6.51 8.28 7.26 6.0 4.5 Wubetie et al., 2018b Eth. 1377/B.brizantha ecotype 2.65 5.72 6.89 6.28 4.7 4.3 Wubetie et al., 2018b

27

2.7. Effect of fertilizer, cultivars and soil type on forage quality

2.7.1. Dry matter content

The dry matter (DM) content of forages is the proportion of total components such as fibers, proteins, ash, water-soluble carbohydrates; lipids, etc. remaining after the water has removed. Dry matter is the percentage of the forage that is not water. Dry matter is also very important, as the moisture content will give clues as to how forage will preserve when stored by baling or ensiling (Schroeder, 2012). The lower the dry matter content, the higher the fresh weight of forage needed to achieve a target nutrient intake, whether this is grazed grass or conserved forage (Brandi, 2015).

According to Biniyame Mihret et al. (2018) reported that type of fertilizer and level had no significant effect on the dry matter content of desho grass. Similarly, Abdi Hassan et al. (2015) on the effects of nitrogen fertilizer application on the nutritive value of Cenchrus ciliaris and Panicum maximum grown under irrigation at Gode, , Ethiopia, reported that fertilizer level and species of grass had no significant effect on dry matter content. Types of cultivars or ecotypes could affect dry matter content of grass. Wubetie Adnew et al. (2018b) reported that dry matter content was significantly affected by the type of ecotypes Brachiaria brizantha grass. The highest dry matter content was recorded from Eth. 13809 (37.75%) followed by Eth.13727 (36.71%) and Eth.1377 (35.71%). The same author reported that agro ecology had a significant effect on the dry matter content of Brachiaria brizantha ecotypes. The highest dry matter content was reported at lowland (37.77%) followed by midland (36.61%) and highland (35.79%). This is due to agroecolgy and harvesting age differences. The information on the effect of soil on the dry matter content is limited in the different countries including, Ethiopia.

2.7.2. Ash and Organic matter

Ash is the inorganic residue remaining after the water and organic matter has been removed by heating in the presence of oxidizing agents, which provides a measure of the total amount of minerals within forage. By nature, ash or minerals are devoid of protein, calories, energy or nutrients that the ruminant can ferment in their rumen. The reason we measure the ash content of forages is to estimate energy and calculate non-fiber carbohydrate content (Patrick, 2005). Organic matter was calculated as ash content subtracted from 100 (100-ash 28

content) dry samples. Ash content showed a decrease as the cutting interval of King Napier grass increased (Lounglawan et al., 2013).

The type of fertilizer and level fertilizer had a significant effect on the ash and organic matter content of forage grasses. Biniyame Mihret et al. (2018) reported that the type of fertilizer had a significant effect on the ash and organic matter content of desho grass. The highest ash content was recorded at manure (17.77%) followed by without fertilizer (16.41%) and NPS fertilizer was the lowest ash content (15.76%) while the highest organic matter content was recorded at NPS followed by manure and the lowest was at without fertilizer. Abdi Hassan et al.( 2015) reported that ash content and organic matter was significantly affected by nitrogen fertilizer levels the level of fertilizer increases from (0, 50 and 100 kg/ha) the ash content was decreased from (14.43 to 11.85 to 11.89 %) respectively, while the organic matter content was increased as the level of fertilizer level increased .

On the other hand, agro ecology and type of acesstions had effect on the ash content of Brachiaria brizantha ecotypes as reported by (Wubetie Adnew et al., 2018b). The highest ash and lowest organic matter content was recorded from lowland (12.22 and 87.88 %) followed by highland (12.17 and 87.83%) and midland (10, 09 and 89.91%) respectively. However, the same author reported that there was a significant difference among the three Brachiaria brizantha ecotypes (Eth. 13726, Eth.13809 and Eth.1377) at all agroecologies and harvesting ages. According to Getnet Assefa and Inger Ledin (2001), the soil type had no significant effect on the Ash and organic matter content of Oat cultivars in the highland parts of Ethiopia at Holeta district. Therefore it is limited information regarding the effect of soil type on these parameters of Brachiaria grass cultivars or related grass families in different countries including, Ethiopia.

2.7.3. Crude protein content and crude protein yield

Protein commonly measured as crude protein (CP), which is 6.25 times the nitrogen content of the forage. Crude protein is used because rumen microbes can convert non-protein nitrogen to microbial protein, which can then be used by the animal. However, this value should be used with some care, as it does not apply to non-ruminants or when high levels of nitrate are present in the forage (Ball et al., 2001).

29

The CP content of desho grass was significantly affected by fertilizer type and level of fertilizer (Biniyame Mihret et al., 2018). The highest CP content was observed at NPS fertilizer (10.95 %) followed by manure application (10.04 %) while the lowest observed at without fertilizer (9 %). Similarly, Abdi Hassan et al. (2015) reported that CP content of Cenchrus ciliaris and Panicum maximum grasses were increased as the nitrogen level of fertilizer increase from 0 kg/ha ( 13.54 %) to 50 kg/ha (17.34 %) to 100 kg/ha (19.84) under irrigation condition.

The variation of CP content can be affected by agro ecology and species or cultivars of forage grass. Wubetie Adnew et al. (2018) reported that overall CP of Brachiaria Brizantha grass ecotypes had higher at midlands (13.35%) followed by lowland (12.5%) and high land (9.56%). The same author reported that numerically higher CP content had recorded at Eth.13809 (11.35%) followed by Eth.13726 (11.23%) and Eth.1377 (10.94%). Similarly, Mustaring et al. (2014) reported that CP content of Brachiaria grass was affected by type and species of grass. The lowest CP was from Brachiaria mutica grass (8.64 %) followed by Brachiaria brizantha (11.77%) while the highest was from Brachiaria Mulato (11.93 %) at similar harvesting age (60 days) in lands of central Sulawesi, Indonesia. Mutimura et al. (2017) reported that the type of species or cultivars of Brachiaria grass affected CP content. The highest CP content was recorded from Brachiaria decumbens (16.7%) and the lowest was from Brachiaria hybrid Mulato-I (13.8%) at 90 days of harvesting age. Susan et al. (2015) reported also the type of species and cultivars had a significant effect on the CP content of Brachiaria grasses. The highest CP was recorded from Brachiaria Marandu (9.2%) followed by Brachiaria La Liberated (7%) and Brachiaria decumbens (4.9) at 60 days of harvesting age.

According to Getnet Assefa and Inger Ledin (2001), the soil type had a significant effect on the CP content of Oat cultivars in the highland parts of Ethiopia at Holeta district. The CP from red soil (7.5%) was higher than black soil (5.9%). This shows that there is limited information regarding the effect of soil type on the CP content of Brachiaria grass cultivars or related grass families in different regions including, Ethiopia.

The CPY content of desho grass was significantly affected by the fertilizer type (Biniyame Mihret et al., 2018). The highest CPY content was observed at NPS fertilizer (2.76) followed by manure application (1.39) while the lowest observed at without fertilizer (0.7) ton per

30

hectare. Type of cultivars or species and agro ecology had a significant effect on the CPY of forage grass. Wubetie Adnew et al. (2018b) reported that the highest CPY of Brachiaria brizantha grass was recorded from Eth.13809 (0.66 t/ha) followed by Eth.1377 (0.52 t/ha) and Eth.13726 (0.38 t/ha). The same author reported that the highest CPY was recorded from lowland (0.71 t/ha) followed by midland (0.52 t/ha) and highland (0.35 t/ha). Regarding the effect of soil, type CPY is limited information with the same or related grass families in different regions as well as in Ethiopia.

On the other hand, harvesting age had a significant effect on the crude protein yield of forage grass. Brachiaria mutica grasses and Brachiaria brizantha ecotypes were affected by the harvesting stage. CPY content was increased with the advancing maturity of forages from 60 to 90 to 120 (Mimilia Zemene, 2018; Wubetie Adnew et al., 2018b) respectively. The effect of days of harvesting indicated an increasing trend in CPY of desho grass with extended days of harvesting. Crude protein yields (CPY) increased progressively and significantly (P<0.01) as growth period increased (0.76 t/ha at 75 days to 2.36 t/ha at 135 days (Genet Tilahun et al., 2017).

2.7.4. Neutral detergent fiber

Neutral detergent fiber is a good indicator of fiber contents in forages, they do not measure how digestible that fiber is, but NDF is a good indicator of "bulk" and thus feed intake. Neutral detergent fiber had shown to negatively correlated with dry matter intake. In other words, as the NDF in forages increases, animals will be able to consume less forage. A better prediction of forage intake can be made using NDF; therefore, better rations can be formulated (Schroeder, 2012). Generally, the NDF content in the feed is one of the major criteria to predict DM intake (DMI) in animals, especially for grazing animals. This is because high NDF content in a feed leads the animal to eat less feed (Lardner et al., 2015) and hence affects animal productivity. Kaplan et al. (2014) showed that cell wall contents increased with increasing maturity. It also reported that temperature in growing areas has a significant effect on the NDF content of forages. Even at the same maturity, NDF contents of forages will be high when growth takes place at high temperature, compared with cool temperature (Linn and Martin, 1999). Feed intake of dairy cattle decreased with increasing NDF content of diets ranging from 22.5% to 45.8% (Arelovich et al., 2008)

31

The type of fertilizer and level of fertilizer could affect the NDF content of forage grass. The NDF content of desho grass was affected by fertilizer type and level of fertilizer (Biniyame Mihret et al., 2018). The highest NDF was observed without fertilizer (61%) followed by manure application (55%) while the lowest observed at NPS fertilizer (51%). Similarly overall NDF content Cenchrus ciliaris and Panicum maximum grasses were decreased as the nitrogen level of fertilizer increases from 0 kg/ha (70.89%) to 50 kg/ha (68.52%) to 100 kg/ha (61.03) as reported by (Abdi Hassan et al., 2015) under irrigation condition.

The type of species or cultivars had a significant effect on the NDF content of Brachiaria grasses. Mustaring et al. (2014) reported that the highest NDF content was recorded from Brachiaria mutica (72%) followed by Brachiaria brizantha (65%) and Brachiaria hybrid Mulato-II (64%) at 60 days of harvesting age. Similarly, Mutimura et al. (2017) reported that the NDF content of Brachiaria decumbens (52.3%) is higher than Brachiaria hybrid mulato (42.9 %) at 90 days of harvesting age. Susan et al. (2015) reported that the higher NDF content was recorded from Brachiaria decumbens (71.3%) followed by Brachiaria brizantha La Liberated (69.1%) and while lowest was recorded from Brachiaria Marandu (65.6 %) and Brachiaria hybrid Mulato-II (63.3%) at similar 90 days of harvesting age. On the other hand, Wubetie Adnew et al. (2018b) reported on ecotypes of Brachiaria brizantha and agro ecology had no significant effect on the NDF content brizantha ecotypes. According to Getnet Assefa and Inger Ledin (2001), the soil type had no significant effect on the NDF content of Oat cultivars in the highland parts of Ethiopia at Holeta district. Therefore, there is limited information regarding the effect of soil type on this parameter of Brachiaria grass cultivars or similar grass families in different countries including Ethiopia.

2.7.5. Acid detergent fiber

Acid detergent fiber is the portion of the forage that remains after treatment with detergent under acid conditions. Acid detergent fiber is a good indicator of digestibility and thus energy intake (Van Soest, 1994). Acid detergent fiber is important because it had shown to negatively correlated with how digestible forage maybe when fed. As the ADF increases, the forage becomes less digestible (Schroeder, 2012). The ADF content of desho grass was significantly affected by fertilizer type and level of fertilizer as reported by (Biniyame Mihret et al., 2018). The highest ADF was observed without fertilizer (52%) followed by

32

manure application (44%) while the lowest observed at NPS fertilizer (41%). Similarly, overall ADF content Cenchrus ciliaris and Panicum maximum grasses were decreased as the nitrogen level of fertilizer increases from 0 kg/ha (50.37%) to 50 kg/ha (44.91%) to 100 kg/ha (42.89%) (Abdi Hassan et al., 2015) under irrigation condition.

Type of species or cultivars had a significant effect on the ADF content of Brachiaria grasses. Mustaring et al. (2014) reported that the highest ADF content was recorded from Brachiaria Mutica (46%) followed by Brachiaria hybrid Mulato-II (39%) and Brachiaria brizantha (38%) at 60 days of harvesting age. Similarly, Susan et al. (2015) reported that the higher ADF content corded from Brachiaria decumbens (42.9%) followed by B. brizantha La Liberated (42.2%) and while lowest was from Brachiaria Marandu (38%) and Brachiaria hybrid Mulato-II (37.5%) at similar 90 days of harvesting age. On the other hand, Wubetie Adnew et al. (2018b) reported on ecotypes of Brachiaria brizantha significant effect on the ADF content brizantha ecotypes. The highest overall mean ADF content was recorded from Eth.13726 (50.77%) followed by Eth.1377 (46.81%) and Eth.13809 (43.47%). Regarding soil type effect on the ADF content of forage grass was limited information on Brachiaria grass cultivars or related grass families in different regions including Ethiopia.

2.7.6. Acid detergent lignin

Lignin is the prime factor influencing the digestibility of plant cell wall material. As concentrations of lignin increase, digestibility, intake and animal performance usually decreases (Chaves et al., 2002). Lignin limits cell wall (fiber) digestion by providing a physical barrier to microbial attack and the concentration of both fiber and lignin increases as plants mature (Van Soest, 1994). Ruminants can digest the cellulose and hemicellulose components of fiber but the lignin inhibits the rate and extent of digestion especially when the proportion of lignin in fiber begins to increase. Lignin precursors also have anti- microbial properties (Chaves et al., 2002). However, it benefits plants by providing mechanical support, water impermeability and protection from insects. Lignin is more prevalent in grass stem than leaf and absent from legume leaves. Therefore, forages with lower ADL concentrations are more desirable (McDonald et al., 2002; Tessema Zewdu et al., 2002) that showed ADL content to increase with the advancing plant age.

33

ADL content of grass species was affected by type and level of fertilizer. The ADL content of desho grass was significantly affected by fertilizer type (Biniyame Mihret et al., 2018). The highest ADL was observed without fertilizer (13.5 %) followed by manure application (9.8 %) while the lowest observed at NPS fertilizer (8.3 %). Similarly Abdi Hassan et al. (2015) reported that the overall ADL content Cenchrus ciliaris and Panicum maximum grasses were decreased as the nitrogen level of fertilizer increase from 0 kg/ha (10.25%) to 50 kg/ha (9.3%) to 100 kg/ha (6.59%) under irrigation condition. Type of species or cultivars had a significant effect on the ADL content of Brachiaria grasses. Mustaring et al. (2014) reported that the highest ADL content was from Brachiaria mutica (11%) followed by Brachiaria hybrid Mulato-II (8%) and Brachiaria brizantha (8%) at 60 days of harvesting age. Similarly, Susan et al. (2015) reported that the higher ADL content was recorded from Brachiaria hybrid Mulato-II (6.4%) followed Brachiaria decumbens (4.9%) and lowest was recorded from Brachiaria brizantha La Liberated (3.3%) from Brachiaria Marandu (3.6 %) at similar 90 days of harvesting age. On the other hand, Wubetie Adnew et al. (2018b) reported on ecotypes of Brachiaria brizantha significant effect on the ADL content brizantha ecotypes. The highest overall mean ADL content was recorded from Eth.13726 (15.3%) followed by Eth.1377 (13.31%) and Eth.13809 (11.3%). Regarding soil type effect on the ADL content of forage grass was limited information on Brachiaria grass cultivars or related grass families in different countries including Ethiopia.

2.7.7. In vitro dry matter digestibility

The in vitro techniques had developed to overcome the shortcoming of the in vivo technique. The advantages of the in vitro techniques are that they are less laborious and are more suitable for a large-scale evaluation of ruminant feeds. In vitro dry matter digestibility of forage grass was affected by the type of species or cultivars of forage grass. Mustaring et al. (2014) reported that IVDMD was affected by the type of Brachiaria species. The highest IVDMD was recorded from Brachiaria brizantha (57%), Brachiaria Mulato-I (56%) and Brachiaria mutica (50%) at 60 days of harvesting age. Similarly, Mutimura et al. (2017) reported that IVDMD affected by the type of species and cultivars. The highest IVDMD was from Brachiaria hybrid Mulato-II (45.7%), Brachiaria hybrid Mulato-I (49.5%) and Brachiaria Marandu (46.5%) at 90 days of harvesting. Susan et al. (2015) also reported that IVDMD affected by species of Brachiaria grass. The highest IVDMD was obtained from Mulato-II (57.5 and 51.4%), B.decumbens (53.8 and 50.1%), 34

Brachiaria La Liberated (54.4 and 43.6%) and Brachiaria Marandu (52 and 28.8%) at 60 and 90 days of harvesting age respectively. Regarding soil type effect on the IVDMD of forage grass was limited information on Brachiaria grass cultivars or related grass families in different countries including Ethiopia.

2.7.8. In vitro organic matter digestibility

The IOMD was affected by Brachiaria species type and species of grass (Mustaring et al., 2014). The lowest IOMD was from Brachiaria mutica grass (47.36 %) followed by Brachiaria brizantha (53.23%) while the highest was recorded from hybrid Mulato-I (58.43 %) at similar harvesting age (60 days) in lands of central Sulawesi, Indonesia. Similarly, Mutimura et al. (2017) reported that types of Brachiaria species and cultivars significantly affected that in vitro organic matter digestibility. Organic matter digestibility content differed (P<0.001) among grass genotypes. The highest IVOMD was from hybrid Mulato- II (50.9), hybrid Mulato-I (49.4) and Brachiaria Marandu (43.7) at 90 days of harvesting age. Getnet Assefa and Inger Ledin, 2001 reported the effect of soil type on IOMD on the Oat cultivars. The highest IOMD was from black soil (56.6 %) than red soil (54.7%). The soil type effect on the IVDMD of Brachiaria grass cultivars grass or related grass families was limited information in different countries including Ethiopia.

2.7.9. Metabolizable energy

Types of Brachiaria grass cultivars significantly affected the ME. Mutimura et al. (2017) reported that ME MJ/kg of B. hybrid Mulato-I (8), B. La Liberated (8.1), B.decumbens (8.4), B.Mulato-II (7.7) and B. Marandu (6.5) at 90 harvesting days. On the other hand, Bimrew Asmare et al. (2017) reported that ME desho grass was not significantly affected by harvesting day but linearly as harvesting increases the ME was decreased from 90 (6.48 to 120 (6.19) to 150 ( 5.87) harvesting days. Regarding soil type effect on the ME of forage grass species was limited information on Brachiaria grass cultivars or related grass families in Ethiopia.

35

Table 2. 6 Summary on Quantity and Quality Brachiaria grass species

B. Species HD ALT DM (%) OM (%) ASH (%) CP (%) CPY Source (Ton/ha B.brizantha 8 Mid 20 86 14 11.77 - Mustaring et al., 2014 B.mutica 8 Mid 22 88 12 8.64 - Mustaring et al., 2014 B.mutica 8 Mid 94 85.4 14.6 13.51 0.7 Mimila, 2018 B.mutica 12 Mid 94 85.9 14.5 9.46 1.17 Mimila, 2018 B.mutica 16 Mid 95 84.9 10.5 6.16 1.27 Mimila, 2018 B.Mulato-I 8 Mid 21 85 15 11.93 - Mustaring et al., 2014 B.Mulato-I 8 Mid 21 87.1 12.9 17.3 - Mutimura et al., 2017 B.Mulato-I 12 Mid 20 88.2 11.8 13.8 - Mutimura et al., 2017 B.Mulato-II 8 Mid - 85 15 10.7 - Susan et al., 2015 B.Mulato-II 12 Mid - 88.6 11.4 7.0 - Susan et al., 2015 B.Mulato-II 12 Low 99 84.2 15.8 13.3 - Clara, 2013 B.Mulato-II - low 90.1 85.48 14.92 10.65 1.41 NegasuYibarkew,2018 B.Marandu 8 Mid - 86.1 13.9 9.2 - Susan et al., 2015 B.Marandu 12 Mid - 88 12 6.2 - Susan et al., 2015 B.La Liberated 8 Mid - 87.5 12.5 8 - Susan et al., 2015 B.La Liberated 12 Mid - 90.4 9.6 7 - Susan et al., 2015 B.decumbens 8 Mid - 87.9 12.1 8 - Susan et al., 2015 B.decumbens 12 Mid - 91.5 8.5 4.9 - Susan et al., 2015 B.decumbens 8 Mid 13 87.3 12.7 18.2 - Mutimura et al.,2017 B.decumbens 12 Mid 32 90.7 9.3 16.7 - Mutimura et al.,2017 B.decumbens 16 Mid 26 90.1 9.9 11.2 - Mutimura et al.,2017 Eth.13726/ B.brizantha 12 Mid 37 90.3 9.66 12.36 0.47 Wubetie et al., 2018b Eth.13809/Bbrizantha 12 Mid 38 90.9 9.15 10.54 0.57 Wubetie et al., 2018b Eth.1377/B.brizantha 12 Mid 35 88.5 11.5 11.17 0.47 Wubetie et al., 2018b

36

Table 2. 7 Summary on fiber contents and digestibility of Brachiaria grass species by different authors

B. Species HD (week) ALT NDF (%) ADF (%) ADL (%) DMD (%) Source B.brizantha 8 Mid 65 38 8 55 Mustaring et al., 2014 B.mutica 8 Mid 72 46 11 47 Mustaring et al., 2014 B.mutica 8 Mid 68.6 34.18 4.17 - Mimila, 2018 B.mutica 12 Mid 67.2 36.13 4.11 - Mimila, 2018 B.mutica 16 Mid 71 36.24 4.43 - Mimila, 2018 B.Mulato-I 8 Mid 39.8 - - - Mutimura et al.,2017 B.Mulato-I 12 Mid 42.9 - - - Mutimura et al.,2017 B.Mulato-I 8 Mid 64 39 8 55 Mustaring et al., 2014 B.Mulato-II 8 Mid 60.6 36.9 2.6 58.6 Susan et al., 2015 B.Mulato-II 12 Mid 63.3 37.5 6.4 53.3 Susan et al., 2015 Mulato-II 12 low 81.2 30.1 - - Clara, 2013 Mulato-II - Low 61.91 31.26 4.2 - Negasu Yibarkew ,2018 B.Marandu 8 Mid 65.6 38.6 2.3 54.3 Susan et al., 2015 B.Marandu 12 Mid 65.6 38 3.6 46 Susan et al., 2015 B.La Liberated 8 Mid 64.7 38.8 3.3 50.5 Susan et al., 2015 B.La Liberated 12 Mid 69.1 42.9 4.4 46.4 Susan et al., 2015 B.decumbens 8 Mid 68.1 42.2 3.4 56.1 Susan et al., 2015 B.decumbens 12 Mid 71.3 42.4 4.9 52.3 Susan et al., 2015 B.decumbens 8 Mid 41.7 - - 41.1 Mutimura et al.,2017 B.decumbens 12 Mid 52.3 - - 52.3 Mutimura et al.,2017 B.decumbens 16 Mid 35.2 - - 35.2 Mutimura et al.,2017 Eth.13726/ B.brizantha 12 Mid 67.5 51.02 16.6 - Wubetie et al., 2018b Eth.13809/B.brizantha 12 Mid 66.9 44.96 12 - Wubetie et al., 2018b Eth. 1377/B.brizantha 12 Mid 63.8 43.07 12.9 - Wubetie et al., 2018b

37

2.8. Farmers’ perception on the varietes or cultivars selection

The inclusion of farmers in the variety selection and perception enabled them to select the best variety/cultivar and/or hybrid of the new forage according to their experiences. Farmers were able to select varieties, which performed well in their local environments. Farmers identified the criteria used for selection through focus group discussions themselves. Like drought tolerance, soil erosion control, plant height, growth habit, the color of leaves, disease and pest tolerance, suitability for grazing and cut-and-carry were used for selection criteria to select different Brachiaria grass cultivars as reported by (Mutimura and Everson, 2012). Similarly, Nkongolo et al. (2008) reported that participatory variety selection (PVS) helps the farmers to select better technologies by comparison to the indigenous forage used by farmers. The same author reported that the criteria chosen by farmers in both districts to select the Brachiaria grasses were similar and the most important were: palatability, high biomass production, drought tolerance, easy to cut, acidic soil tolerance and regrowth capacity. This relationship between farmers’ indigenous knowledge on selecting new forages and their chemical composition supports Abebe Mikoya et al. (2008) who found that nutritive values of fodder from laboratory analyses in two districts of Ethiopia corresponded with the ranks of those feeds given by farmers. The participatory evaluation by farmers in the trial is important for the adoption of new forage technology and its expansion to other smallholder farmers (Mutimura and Everson, 2012).

Variety selection had usually conducted through a top-to-down approach based on agronomic data and chemical composition. Such a type of selection approach is necessary but not sufficient for variety selection. Therefore, participating farmers' perception of variety selection is not a choice rather it is necessary to enhance the adoption of the specific variety. An individual participant farmer has its preference participatory selection of different forage species like lablab (Beyadglign Hunegnaw et al., 2016), cowpea (Misgnaw Walie et al., 2016) and sweet lupin (Wodimeneh Mekonnen et al., 2016) varieties intercropped on maize crop production. The pair wise ranking method is used to prioritize the selection parameters) and the preference-ranking method has employed for variety selection as reported by the same author. Those methods are emphasizing on why an individual farmer selects specific forage varieties

38

Chapter 3. MATERIALS AND METHODS

3.1. Description of the study areas

The study was conducted simultaneously at Andassa livestock research center (Medabit forage trail site) and North Mecha district (Ambomesk kebele) of West Gojam Zone, , Ethiopia. Andassa livestock research center (ALRC) is located 21 Km far from Bahir Dar, Capital of Amhara region and 486 km far from Addis Ababa, of Ethiopia. The center is located 11°29°N latitude and 37°29o E longitude and 1730 meters above sea level. Annual rainfall, maximum and minimum annual temperature was 1330.4 mm, 27.9°c and 13°c respectively (ARMA, 2018). The soil in the area is black clay (vertisols) (Wondimeneh Mekonnen et al., 2016). While Mecha district is 34 km far from Amhara Regional state , Bahir Dar and 520 km far from the national capital city, Addis Ababa. The district is located 9° 23’ to 9° 26’ N latitude and 41° 59’ to 42° 02’E longitude and 1807 -2300 m above sea level. Annual rainfall, maximum and minimum annual temperature was 3043.9 mm, 28.01°c and 10.57°c respectively (ARMA, 2018). The soil in the district is dominantly is red (nitosoil) (Likawent Yeheyis et al., 2012). The pattern of rainfall in both districts is monomodal.

Figure 3. 1 Map of study districts 39

3.2. Soil sampling and chemical composition analysis

From both soil types, soil samples were collected in May before planting of experimental Brachiaria grass cultivars at depth 30 cm using an auger for analysis. Plant litter and other dirties on the soil surface were removed before collecting the samples. Soil samples collection was followed diagonal (X) method of sampling and composite soil samples were made and put into a sterile plastic bag. The collected soil samples were mixed and taken as one composite sample and brought to Andassa livestock research center, air-dried in an open space to make it dry and then after drying the coarse particles of the soil have been crushed with mortar and pestle to make fine and easy of grinding with the mill. After drying and grinding the two soil samples per block totally, six samples were sent to Adet Agricultural Research center (AARC) for chemical analysis and pH determination of experimental sites. The two mm sieve soil sample size was used for soil chemical compositions analyses.

The collected soil samples from both soil types were analyzed for soil pH, organic carbon (OC), total nitrogen (N), available phosphorus (P) and organic matter (OM) using standard laboratory procedures at AARC in soil laboratory. Total N in the soil was determined by the Kjeldahl method (Dewis and Freitas, 1975). According to Walkley and Black (1934), procedure organic carbon content in the soil was determined by the reduction of potassium dichromate by organic carbon compound and determined by reduction of potassium dichromate by oxidation-reduction titration with ferrous ammonium sulfate. Available P in the soil was determined by Olsen's method using a spectrophotometer (Olsen et al., 1954). The pH of the soil using 1:2.5 (weight volume-1) soil samples to CaCl2 solution ratio was determined with a glass electrode attached to a digital pH meter (FAO, 2008). Finally, organic matter was calculated as (OC*1.72) standard methods as described by Okalebo et al. (2002).

The current result on the chemical composition of soil samples of experimental sites before the experiment were described in (Table 3.1). The analysis result for pH, organic matter and available phosphorus were significantly (P< 0.05) different among the black and red soil types of experimental sites. The higher pH content (6.94) was observed at black soil than red soil (5.45). Moreover, the higher organic matter (2.86) was recorded at red soil than black soil (2.42). Similarly, the highest available phosphorus content (7.49) was observed

40

at red soil than black soil (4.57). The organic carbon and total nitrogen contents were not significantly different (P< 0.05) between two soil types.

According to the Landon (1991), classification of soils having total nitrogen greater than 1.0% as very high, 0.5-1.0% high, 0.2-0.5% medium, 0.1-0.2% low and less than 0.1% as very low in total nitrogen content. Similarly Tekalign Tadesse (1991) classified soils having total N availability of < 0.1 % as very low,0.1-0.2% low,0.2-0.5 medium,0.5-1% high and >1% very high. Therefore, the overall mean of soil sample chemical analysis showed that in the current study of both experimental sites (red ,0.23) and (black,0.22) soils had medium mean of total nitrogen content (0.225%) as reported by both authors. The chemical analysis soil showed a pH of 6.94 and 5.45 at black and red soil respectively. Tekalign Tadesse (1991) reported that guidelines for soil pH values as pH > 9.0 (very strongly alkaline), 9.0– 8.5 (strongly alkaline), 8.4–7.9 (moderately alkaline), 7.8–7.4 (mildly alkaline), 7.3–6.6 (neutral), 6.5–6.1 (slightly acid), 6.0–5.6 (moderately acid), 5.5–5.1 (strongly acid1), 5.5–5 (strongly acid), 5.0–4.5 (very strongly acid) level of classification. Thus, the chemical reaction (pH) of the experimental site was neutral and strongly acid for black and red soil respectively (Table 4.1). However the result of both soils were in the range of PH (4.2 to 8) that Brachiaria grass species give production (Larry, 2013).

The current result showed that the total available phosphorus level in the experimental sites was 4.57 and 4. 49 mg/kg for black and red soil respectively. Tekalign Tadesse (1991) described soils with having available P <10, 11-31, 32-56, >56 ppm as low, medium, high and very high, respectively. Therefore, the result showed that the total amount of available soil P level in the current study was classified as low. The current result showed that the organic carbon content of experimental sites was 1.41 and 1.64 % for black and red soil respectively. Hazelton and Murphy (2007) classified soil organic carbon percentages of < 0.6, 06-1, 1-1.8, 1.8-3 and > 3 as very low, low, medium, high and very high, respectively. Thus, the current result of soil organic carbon content at both soil types was medium level classification as reported by the same author.

41

Table 3. 1 Soil sample chemical composition of experiment sites before planting

Soil N Soil chemical composition parameters type (N=12) pH Organic Organic Total nitrogen Available carbon (%) matter (%) (%) phosphorus (Ppm) Black 6 6.94a 1.41 2.42 b 0.22 4.57b Red 6 5.45b 1.64 2.86a 0.23 7.49a Mean 6.194 1.52 2.64 0.225 6.03 SEM 0.22 0.06 0.11 0.01 0.43 Sig *** Ns ** Ns ** CV 0.908 12.876 12.188 17.117 10.4

Ppm=part per million, SEM=standard error of the mean, CV=coefficient of variation, Ns= not significant, Sig= significance level

3.3. Experimental land preparation

The two districts (Bahir Dar Zuria and Mecha) were selected purposely based on their soil type history and accessibility, Andassa kebele from Bahir Dar Zuria district and Ambomesk kebele from Mecha districts were selected as experimental sites. From each experimental kebeles, the appropriate site was selected by visiting different sites by considering waterlogged areas, which are not favorable for most Brachiaria grass cultivars. After selection, the land was cleared, ploughed by tractor and harrowed again traditional oxen on 20 to 30 days before laying out plots and planting a fine tithe to facilitate soil aeration, to dry and removed or reduced unwanted weeds.

3.4. Treatments and Experimental design

The study was conducted at red and black soil type simultaneously, of which, all- experimental materials, design and treatments were the same across two soil types. The experimental design used in the current study was a factorial arrangement of treatments in a randomized complete block design (RCBD) consisting of three factors (fertilizer level, soil type and cultivars) with three replications. The experiment consisted of six Brachiaria grass cultivars (mutica, hybrid Mulato-I, hybrid Mulato-II, Marandu, La Liberated and decumbens), two soil types (black and red) and two levels of fertilizer (with and without fertilizer) at begining of experiment.The experiment had 24 treatments with a (2*2*6) 42

factorial combination of two levels of fertilizer, two soil types and six Brachiaria grass cultivars. There were three blocks, each containing 12 plots resulted in 36 plots in total with each plot measuring 3m x 3 m at both soil types. The inter-row and intra-row spacing were the same for all treatment 0.5m. The spacing between block and plots was 1 and 0.5 meters respectively. The total area of the experimental land was 504m² (42m *12m). The plot size of each treatment was 9m2 (3mx3m) and the net harvestable plot area was 6m² (3m*2m) used by excluding one outer row on both sides of each plot. The experimental materials (cultivars) of root splits were collected from Andassa livestock research centers (ALRC) and Woreta national rice research center (WNRRC). After proper land preparation, the cultivars were planted at two locations in June by root split propagation mechanism in the main rain season. The fertilizer was applied at the time of planting, with a rate of NPS 100kg/ha and urea with a rate of 50kg/ha was applied after 30 days of planting grass (Abdi Hassan et al., 2015).

According to the recommendation of Cameron et al. (2008), two-thirds of the Brachiaria grasses root splits were buried at the depth of 10-15 cm on a well-prepared plot and the apical third was left on the ground. Weeds were controlled by hand weeding to avoid interference by interspecific competition and early and then three times per month until the final harvesting was accomplished to eliminate regrowth of undesirable plants and removal of the dry root to promote fodder re-growth by increasing soil aeration. To ensure the trial was not be destroyed by wild or domestic animals, the entire experimental area was fenced off and kept by assigned enumerators. The data from Brachiaria decumbens cultivar was not collected due to survival problems at both soil (no survive at red and less than 6% at black soil), so this cultivar was removed from the presentation and discussion part. Therefore, only 20 treatments (2*2*5) were presented throughout all periods. The survival rate of each experimental cultivars was listed at the appendix table (Appendix 7.1.5). In my observation, this survival problem of Brachiaria decumbens in the current study could be low acidic tolerance ability and the presence of very thin roots.

3.5. Data collection

Data on all morphological characteristics and yield parameters were collected at 90 days harvesting age to have optimum forage nutritive value and dry matter yield of Brachiaria grass (Mimilia Zemene, 2018). In each plot, ten plants were randomly selected by avoiding 43

border plants to a record number of tiller per plant (NTPP), a total number of leaves per plant (NLPP), leaf length per plant (LLPP), length and number of root per plant and leaf: stem ratio (LSR). Sample tillers from each randomly taken plants were used to determine the number of leaves per tiller (NLPT) (Tessema Zewdu et al., 2003). Harvesting was done by hand using a sickle, leaving a stubble height of 10 cm recommended by Tudsri et al. (2002). Farmers' perception was done before the harvesting stage of Brachiaria grass at 90 days. For partial budget analysis, all costs and benefits were collected at each step. Morphological parameters such as plant height, leaf length, and root length were measured using measuring tape. Number of tillers, number of leaves and number of roots per plant was computed by counting. Counts took from ten plants that were randomly selected from the middle of rows by excluding the two border rows of each plot at each soil type at 90 days of harvesting after planting.

3.5.1. Plant height

The mean plant height was taken from the net plot area of each soil type for plant height analysis. Plant height measured on the primary shoot from the soil surface to the base of the top-most leaf using a meter rule as described by Rayburn et al. (2007). It measured from base to tip of the upper leaves of the main stem.

3.5.2. Number of tillers, leaves and leaf length per plant

The number tillers per plants of Brachiaria grass cultivars were counted from the randomly selected sample of ten plant of each plot from middle two rows and mean was calculated. Main stem was also included to calculate the total tillers per plant. Number and length of leaves per plant of were counted from the sample of ten plants. Mean leaves per plant were calculated and then the total number of leaves per plant were estimated from the tiller number per plant. Leaf length per plant of was measured by graduated ruler from the base of the collar region of the leaf to the tip of the leaf. Mean of leaf length was taken from randomly selected sample plants from the two middle rows.

3.5.3. Number and length of roots per plant

The number and length of root were counted and measured after the shoot parameters were accomplished. At the end of each harvesting day, ten sets of root samples were collected

44

from all plots with an auger at 0.5m depth. The ten sets of samples from each plot were clean up from any residual soil and other debris. The numbers of roots per plant were counted and the mean was calculated. The length of the root was measured from the crown part to the tip of the root by graduated ruler and their mean was calculated.

3.5.4. Leaf to stem ratio and forage yield

Leaf to stem ratio determined by cutting plants from randomly selected two middle rows, separating into leaves and stems, drying and weighing each component and estimated by dividing leaf dry weight to stem dry weight. The mean of each parameter was calculated (Susan et al., 2016). A fresh herbage yields of Brachiaria grass cultivars were measured immediately after harvesting and weighed on the field soon after mowing using a field sensitive spring balance. One kg Sub-samples taken from each plot at both soil types to determine dry matter content. Leaf and stem dry weight divided by leaf and stem fresh weight and multiplied by 100 to determine DM % for each sample. On the basis of the DM % and fresh biomass yield from the sample area of each plot were used to calculate total dry matter yields for each plot, thereafter, converted to tons per hectare. Finally, adequate quantities of sub-samples were air-dried and stored in airtight bags to be used for chemical analysis. After drying, samples were ground to pass a 1-mm Wiley mill screen at Bahir Dar University College of agriculture and environmental sciences in animal nutrition laboratory and 100-150 gram amount of each was stored in airtight bags for different chemical analyses.

3.6. Forage quality analysis.

Samples at each soil type were subjected to chemical analysis for the determination of organic matter following the methods of (AOAC, 2004). Forage nutritive value measurements such as determination of crude protein by Kjeldahl procedure (AOAC, 2004), acid detergent fiber (ADF), neutral detergent fiber (NDF) and acid detergent lignin (ADL) were analyzed using (Van Soest et al., 1991). Ash was determined by igniting at 550 °C total DM by drying at 105 °C for three hours. In Vitro organic matter digestibility (IVOMD) and In Vitro dry matter digestibility (IVDMD) were determined according to Tilley and Terry (1963). All the chemical analyses were done at the Holeta Agricultural Research Center, Animal Nutrition Laboratory by conventional chemical analysis procedures.

45

The metabolizable energy (ME) was estimated from digestible energy (DE) and in vitro organic matter digestibility (IVOMD), based on the National Research Council (NRC, 2001) formula using the following steps:

DE was obtained using Equation: DE= {0.01 x (OM/100) x (IVOMD +12.9) x 4.4} -0.3, where DE is the digestible energy in calories, OM is the organic matter and IVOMD is the in vitro organic matter digestibility in joules. Then, ME =0.82 x DE (Mcal /kg) was calculated and converted to SI units (MJ/kg) by multiplying by 4.184 (NRC, 2001). Finally the crude protein yield (CPY, t/ha) was determined by the multiplying of DMY by crude protein content (CP) at each soil types. CPY= CP*DMY

3.7. Farmer's perception on experimental Brachiaria grass cultivars

There were 20 and 16 participant farmers in black and red soil, respectively based their interest towards the technologies, willingness to participate on cultivar selection and who had previous experience on the improved forage production. These participant farmers were selected from different community groups. Majority of the farmers were male farmers. The number of female farmers were 6 in black soil site and 4 in red soil. The district agricultural office of forage experts, Andassa livestock research center researchers (feed, nutrition and extension researchers) and kebele development agents (DA) had also taken part in the implementation process of during capturing of farmers perception at filed. After explaining the objectives of study, farmers’ were given the chance to visit all cultivars that are grown in all plots and list or think of their cultivar selection criteria. Then with group discussion they are listed all important selection criterion based on their indigenous knowledge. The selection criterion in both soil type were: plot cover, number of tiller, plant height, leafiness and smoothness.

Pair-wise matrix ranking and preference ranking tools were employed to capture selection criteria’s and select Brachiaria cultivars, respectively. By using the pair-wise matrix ranking method, farmers’ were set the magnitude of each criterion. Finally, the farmers’ were voted by each criterion for the evaluated cultivars with preference ranking method as reported by (De Boef and Thijssen, 2007) a guide on participatory tools working with crops, varieties and seeds. Accordingly, the first rank in selection criteria have a higher magnitude (highest score) while the last ranked criteria were the lowest magnitude (lowest score) to

46

calculate the weighted value of each cultivar at both soil types. A total weighed ranking of Brachiaria cultivars were calculated as the product of the value for the criteria and the score for each specific cultivars. Finally all Brachiaria cultivars were compared with each other based on total or sum of all selection criterion weighed values by preference ranking. Assessing of farmers perception through Brachiaria cultivars selection were done before harvesting age of grass (90 days).

3.8. Partial budget analysis

Information on all variable costs related to fertilizer cost (cost of NPS and Urea) were recorded during the purchasing at cooperatives. The average yield of each treatment was collected, recorded at each soil type. The average purchasing price of NPS per quintal was 1476 birr while the price of Urea was 1250 birr per quintal at cooperatives during our experiment periods. There was no field price or market price for fresh Brachiaria grass cultivars. Therefore, we use average related prices for 1 kg of hay and green grass. The price of one kg Brachiaria grass estimated to Andassa livestock research center balled hay is derived from the price given for balled hay (weighing 18 kg), which has valued 80 Birr accordingly 1 kg hay has been valued 4.44 Birr per kg (ALRC, 2018). Similarly Shimelis Mengistu (2018) reported that 1 kg of hay was 2.06 birr per kg in Damot Gale District of Wolaita Zone, SNNPR. On the other hand price of green forage in the local market of Amhara Region like Debre Tabor and Injebera was estimated to 2.6 birr per kg (personal observation in open market in July 2019 ) .Therefore the average of these prices were assumed to market price estimation of Brachiaria grass cultivars in the current study. Market price for Brachiaria grass (p) = (4.44+2.06+2.6)/3= 3.03 birr per kg. The partial budget analysis (PBA) was employed using the procedure of recommended by (CIMMYT, 1988). Adjusted yield (ADY): is the experimental yield scaled down by a given proportion to approximate the yield that farmers can obtain on their farms. Farmers using the same technologies would obtain yields 10% lower than those obtained by the researchers. Gross benefit (GB): The gross benefit for each treatment was calculated by multiplying estimated market price related feed resources by adjusted dry matter yield. Gross benefit = field price X Adjusted yield Total variable costs (TVC): This is the sum of all the costs that vary for a particular treatment. 47

Net benefit (NB): This was calculated by subtracting the total costs from the gross field benefit for each treatment. NB = GB – TVC. Net benefits are not the same thing as profit, because the partial budget does not include the other costs of production which is not relevant to this particular decision. Dominance analysis (D): This was carried out by first listing the treatments in order of increasing costs that vary or increasing net benefits. Any treatment that has net benefits which are less or equal to those of a treatment with lower costs that vary is dominated. Marginal rate of return (MRR): This was computed by dividing the marginal net benefit (i.e., the change in net benefits) with the marginal cost (i.e., the change in costs) multiplied by hundred and expressed as a percentage. MRR (%) = (∆NR/∆Total Variable Cost)*100, where MRR (marginal rate of return) is a measure of increase in net income that is associated with each cost.

3.9. Methods of Data Analysis

The collected data were managed and organized with MS-Excel (2013). Initially, the dataset was checked for outliers by doing a Shapiro–Wilk’s and Levene’s tests for the analysis of normality of data and homogeneity of variances, respectively. The T-test was employed for the soil chemical composition analysis. All collected biological data were statistically analyzed using the procedure outlined by (Steel and Torrie, 1986) for a factorial experiment in a randomized complete block design using General Linear Model (GLM) procedure of SAS statistical computer package version 9.0 (SAS Institute Inc., 2002) according to Gomez and Gomez (1984). Data on morphological characteristics, total DM yields, and nutritive values were subjected to analysis of variance (ANOVA). Pearson correlation (r) was run to describe the relationships between plant morphology, dry matter yield and forage nutritive value parameters. With the assumptions attended, analysis of variance was performed with application of the F test and for the variables in which the F test was significant, treatment means were compared by Duncan’s multiple range test (DMRT) (p <0.05). Differences were considered statistically significant at a 0.05% significance level. In the all data analysis part and the model the experimental group Brachiaria decumbens, was not covered due to survival problems on both soil types.

48

The statistical model for the analysis of data was: Yijk = μ + Fi + Cj + Sk+ Fi*Ci + Ci*SK +

Fi*SK + Fi*Cj*SK + Eijk

Where; Yijk= all dependent variables (morphological data, forage yield and forage nutritive value) μ = Overall mean

Fi = the effect of fertilizer (with and without fertilizer)

Cj=effect of cultivars (five cultivars)

SK=effect soil type (black and red)

Fi*Ci= interaction effect of fertilizer and cultivars

Ci*SK= interaction effect of cultivars and soil

Fi*SK= interaction effect of fertilizer and soil

Fi*Ci*SK= interaction effect (fertilizer, cultivars, and soil type)

Eijk = residual error

49

CHAPTER 4. RESULTS AND DISCUSSION

4.1. Effects of fertilizer and soil type on morphological characteristics and dry matter yield of Brachiaria grass cultivars

The overall analysis of the current study showed that there were significant interactions between the effects of fertilizer, type of cultivars and soil types.The interaction effects of fertilizer, cultivars and soil type of Brachiaria grasses on plant height, tiller number, number of leaves per plant, leaf length, number and length of root, leaf to stem ratio and dry matter yield were presented in (Table 4.1, Appendix 7.1.3). The interaction effect of fertilizer, type of cultivars and soil type had asignificant effect on morphology and dry matter yield of Brachiaria grass cultivars. All experimental interaction effect parameters were presented separately, so interaction effects fertilizer, soil type and type of cultivars were only presented and discussed throughout the document.

4.1.1. Plant height

The interaction effects between fertilizer, type of cultivars and soil type had significant (P< 0.001) effect on plant height of Brachiaria grasses (Table 4.1). The longest plant height (1.55 and 1.45m) were measured from B.mutica grass at black and red soil with fertilizer respectively, followed by the same cultivar at the black (1.12m) without fertilizer application. The intermediate plant height was measured from the same B.mutica cultivar (0.85m) at red soil and hybrid Mulato-II (0.81m) at black soil without and with fertilizer, respectively. The shortest plant height was recorded from Brachiaria grass cultivars of Marandu (0.32m), La Liberated (0.3m) and B.Mulato-I (0.36m) without fertilizer application at both soil types. Overall, the interaction mean of plant height was (0.62m). This indicates that plant height can be different among different species with the same management system, might be due to the genetic potential of species to extract minirals.

Generally, the interaction effect of fertilizer, cultivars and soil type had a significant (P< 0.001) on plant height. The plant height was increased as increasing levels of fertilizer from zero (0) to 150 kg/ha in both soil types and all-experimental Brachiaria grass cultivars. The overall interaction result showed that the highest plant height recorded from black soil was significantly (P< 0.001) higher than red soil in all experimental cultivars. This might be the acidity level of experimental site of black soil (6.94) was lower than red soil (5.45) and due 49

to the earlier (February) beginning of rainfall distribution at black soil site, this might have created favorable condition for longer plant height at black soil than red soil and might be absorvation and ingestation of soil Corban at red soil during experiment. The plant height increased progressively with increasing fertilizer from 0 to150 kg/ha at both soil type and at all cultivars, associated with massive root development and efficient nutrient uptake, allowing the plant to continue to increase in height of the grass (Berihun Melkie, 2005).

On the other hand.the current result of plant height is in line with findings of to Worku et al. (2017) who reported that the plant height of desho grass increased 0.9 to 1.3 m as an increase of fertilizer level from 100 to 350 kg/ha, respectively. The plant height (1.5m) was measured from Brachiaria mutica through interaction effect of soil, cultivars and fertilizer grass at black soil from the current result was higher than desho grass (1.3m) at 120 days of harvesting stage as reported by the same author. All other experimental cultivars were lower plant height than desho grass as reported by Worku et al. (2017). This difference might be from the genetic differences of grasses, soil type, rainfall and level of fertilizer and harvesting age. Similarly overall mean of current longer and shorter height Brachiaria grass cultivars of current results are higher and linearly similar respectively with desho grass as reported by Biniyame Mihret et al. (2018). The mean plant height of desho grass significantly affected by fertilizer type, of which the longest plant height was from NPS fertilizer (55.23cm) than manure (52.22cm) and without fertilizer application (42.01cm) at similar 120 days of harvesting age. This difference might come from the genetics of grass cultivars, environment conditions, and management practices.

Additionally, the plant height of current result is in agreement with the findings of Nafatul et al. (2015) who noted that plant height increases as the level of NKP fertilizer increased from 0 to 150 to 300 kg/ha. The mean plant height of Brachiaria brizantha Marandu was increasing from 6.74 to 7.86 to 10.13cm respectively at 60 days of harvesting age, which established by the seed propagation system, but the current finding was higher than this study in all experimental Brachiaria grass cultivars. This difference might be coming from a genetic variation of grasses, soil type fertility, rainfall condition, harvesting age and method of the establishment were the experiment was conducted (Campos et al., 2013).

Moreover, effects of type of ecotypes, fertilizer and soil type in the current study were a significant effect on the plant height is in line with Brachiaria brizantha ecotypes as reported

50

by Wubetie Adnew et al. (2018b) who reported on three ecotypes in northwestern Ethiopia. Their finding elucidated that the overall mean of the longest plant height was recorded from Eth. 1377 (58.73cm) followed by Eth. 13809 (51.47cm) and Eth. 13726 (51.07cm).The overall results obtained in the three agro-ecologies: lowland (60.5 cm), midland (57.62 cm) and high land (43.15 cm) indicated that plant height was affected by agro ecology. The current interaction effect result from all experimental cultivars at both soil types except Brachiaria Marandu and La Liberated was longer than as the reports of the same authors. This difference might come from environmental conditions, accession variation and management systems where both studies were conducted. Similarly, the current result is in line with Mustaring et al. (2014) indicated that B.mutica had the highest plant height (207.47cm) than B .brizantha (145.43cm) and B. Mulato-I (117.23cm)at 8 weeks of harvesting. The current result from all experimental cultivars and both soil types were shorter than the same Brachiaria species as reported same author. This difference might come from the accession of grass, soil type, and fertility, management system and harvesting stage where both studies were conducted.

4.1.2. Number of tillers per plant

The result of the current study showed that there were highly a significant (p<0.001) interaction effects of fertilizer, cultivars type and soil type on the number of tiller per plant (NTPP) of Brachiaria grass (Table 4.1). Significantly (p<0.001) the maximum number of tillers per plant were obtained from hybrid Mulato-II (46 and 45.25) with fertilizer at black and red soil respectively, followed by the mutica (40.3 and 41.67) with fertilizer at black and red soil respectively. The intermediate number of tillers per plant were counted from hybrid Mulato-II (39 and 36) with fertilizer at red and black soil, respectively. The minimum number of tillers per plant were obtained from Brachiaria mutica (20.3) without fertilizer at red soil and from La liberated (24.9 and 26.2) without and with fertilizer and Brachiaria Marandu (26.9) without fertilizer at black soil. Overall, interaction means the number of tillers per plant were (34.34). This indicates that application fertilizer, cultivars, and soil type interaction had a significant effect on Brachiaria grass is a critical factor for the number of tillers per plant in the current study.

The effects of fertilizer, cultivars and soil type had a significant (P< 0.001) effect on the number of tillers per plant were increased as increasing level fertilizer from zero (0) to 150

51

kg/ha in the current study at both soil types, in all-experimental Brachiaria grass cultivars. This current result is in line with Worku et al. ( 2017) who indicated that there was a significant difference (p<0.05) in the number of tillers per plant of desho grass as the fertilizer rate increased from 100 to 350 kg/ha which were 54.6 and 79.6 tillers per plant respectively. The number of tillers per plant resulted from all experimental cultivars and the overall mean of the number of tillers per plant in the present study in all experimental cultivars (34.32) were lower than desho grass as reported by Worku et al. (2017). This difference might due to plant type, the difference in plant spacing and harvesting stage, weather, soil type and fertility, level of fertilizer etc. Similarly the current result is also in agreement with Biniyame Mihret et al. (2018) who reported for desho grass that number of tillers per plant of desho grass significantly affected by fertilizer type of which the number of tiller per plant was recorded from NPS fertilizer (61) higher than manure (50) and without fertilizer application (37). This current result of the number of tillers per plant were lower than desho grass as reported by the same author, might be due to fertilizer type, environment conditions, grass type, management system, and harvesting age. Tillers (number) in the current study was different among cultivars and more tiller number recorded at fertilizer treatment groups with the interaction of soil type and type of cultivars. Tiller number is an important attribute of grasses, it increases the chances of survival and the amount of available forage production as reported by (Laidlaw, 2005; Van Saun, 2006). Moreover, it is an indicator of the efficient utilization of all-important resources applied to it by the different grass species. Kizima et al. (2014) reported that the application of the optimal level of Nitrogen fertilizer significantly affects the appearance of new tillers and increases the dynamics of the tiller population of Cenchrus ciliaris.

The number of tillers per plant were affected by the interaction between the type of cultivars, fertilizer and soil type in the current study is in line Wubetie Adnew et al. (2018b) who reported on Brachiaria brizantha grass ecotypes in northwestern Ethiopia. The overall mean of the maximum number of tillers per plant were counted from Eth. 13809 (33.53) followed by Eth. 1377 (29.38) and Eth. 13726 (26.35). The overall number of tillers per plant were obtained in the three agro-ecologies is 33.62, 28.52 and 27.12 for midland, lowland, and highland respectively. The current results of all cultivars except Mutica at both soil types were higher than Brachiaria brizantha ecotypes as reported by the same author. This difference might come from environmental conditions, management system and accesstion

52

variation of grass species besides interaction effects where both experiments were conducted. Additionally current result is in line with Mustaring et al.(2014), who reported that tiller number different among cultivars and increased as maturity, which Brachiaria Mulato-I (122.4) had the highest (P<0.05) tiller number than Brachiaria brizantha (63.73) and Brachiaria mutica (57.7) at similar harvesting age (60 days). This result was significantly higher than from all current results of all experimental groups at both soil types. This difference might be coming from the acesstion variation of grasses, environment conditions, rainfall, soil type, harvesting age, fertilizer, and management practices where both studies were conducted. Similarly the number of tillers per plant were affected by type species or cultivars, fertilizer and soil type in the current study is in line with Susan et al. (2016) reported that mean Llanero (30.5) and Marandu (16.8) was among the highest and lowest tiller number per plant respectively. MG4 (24.5), Piata (25.5), Xaraes (25.5), Mulato- II (23.8) and Basilisk (20.5) also recorded high and similar tiller numbers with Llanero at 16 weeks. The current interaction results from all experimental cultivars were higher than as reported by the same author even with these same cultivars of Brachiaria Marandu and Mulato-II in the current study. This variation might come from accession variation, harvesting age, soil type, establishment methods and fertilizer application (Campos et al., 2013) where both experiments were conducted.

4.1.3. Number of leaves per plant

The current result indicated that there were highly a significant (p<0.001) interaction effects of fertilizer, soil type and type of cultivars on the number of leaves per plant (NLPP) of Brachiaria grasses (Table 4.1). The maximum number of leaves per plant were recorded at hybrid Mulato-II (380.93 and 366.68) with fertilizer at black and red soil, respectively, and significantly higher than at (p<0.001) from all experimental groups, followed by mutica cultivar (333.3) with fertilizer at black soil. The intermediate number of leaves per plant were obtained from mutica (264) and hybrid Mulato-I (244) with fertilizer at red soil and was not significantly different among each other (p<0.001). The minimum number of leaves per plant were obtained from Mutica (135) and Brachiaria La Liberated (124.07) without fertilizer at red and black soil respectively, which significantly lower than from all experimental cultivars (p<0.001). The overall mean of the number of leves per plant obtained from the current study were (212.76).

53

The number of leaves per plant were significantly affected by interaction effects of fertilizer, type of cultivars and soil type all cultivars in the current study is in line with findings of Biniyame Mihret et al. (2018) who reported that number of leaves per plant of desho grass was significantly affected by fertilizer type and level of fertilizer. The same author reported that mean number of leaves per plant were highly decreased from NPS (530 leaves per plant) to manure (388) to without fertilizer (307) at 120 days of harvesting age. The Number of leaves plant form Mulato-II at both soil types and Brachiaria mutica at black soil in the current study is comparable with desho grass as reported by the same author at manure and control groups respectively. The overall mean number of leaves per plant (212.76) in the current study was significantly lower than desho grass as reported by the same author at 120 harvesting age. This difference might come from genetics, soil type and fertility, type of fertilizer, environment condition, altitude, management system, and harvesting age where current study was conducted.

Similarly, the current result number of leaves per plant are in line with Nafatul et al. (2015) reported that as the level of NKP fertilizer increases form (0 to 150 to 300kg/ha) the number of leaves per plant of Brachiaria Marandu increased from (4.87 to 6.62 to 10.12) respectively. The current result from all cultivars at both soil types are significantly higher than the same species Brachiaria Marandu as reported by Nafatul et al. (2015). This difference might come from soil type and fertility, type of fertilizer, environment condition, method of propagation and harvesting age as reported by (Campos et al., 2013) where both experiments were implemented. As increases, the application of fertilizer through soil type and cultivar interaction creates the greater the numbers of leaves, which are important for the photosynthetic and transpiration surface, which, produced from the newly emerging tillers... This could be the addition of the same rates of nitrogen in the experiment, which contributed to an increase in the average number of leaves per plant in some experimental cultivars. This was probably because use nitrogen increased plant growth and plant height which resulted in more nodes, internodes and consequently more leaves per plant (Abdi Hassan et al., 2015).

On the other hand, the number of leaves per plant were affected by the interaction effect type of cultivars, fertilizer and soil type in the current study is in line with Wubetie Adnew et al. (2018b) reports of three Brachiaria brizantha grass ecotypes. The overall mean of the maximum number of leaves per plant were recorded from Eth. 1377 (5.01) followed by Eth. 54

13726 (4.74) and Eth. 13809 (4.36). The overall number of leaves per plant were obtained in the three agro-ecologies were 5.04, 4.96 and 4.11 for midland, lowland, and highland respectively. The current result obtained from all experimental cultivars at both soil types were very higher than as reported by the same author. This difference might be due to accession, soil type and fertility, type of fertilizer, altitude, management system in addition to interaction effects of fertilizer, soil type and cultivars where both studies were conducted (Campos et al., 2013).

4.1.4. Leaf length per plant

The interaction effects between fertilizer, type of cultivars and soil type on the length of leaves per plant (LLPP) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.1). The longest leaf length per plant was recorded from B.mutica (22cm) with fertilizer at black soil significantly different from all others followed by Brachiaria Marandu (20.27cm), hybrid Mulato-II (20.2cm) and Brachiaria La Liberated (20.58cm) with fertilizer at red soil which was not significantly different each other. The intermediate length of leaf per plant was obtained from Mulato-II (17.5cm) without fertilizer, hybrid Mulato-I (17.9cm) with fertilizer, B.mutica (18cm) at black soil and from Brachiaria Marandu (17.13cm) at red soil of which were not significantly different each other (p<0.001). The shortest overall mean length of leaves per plant was obtained from hybrid Mulato-II (10.74cm) and Brachiaria La Liberated (11.17cm) without fertilizer at red and black soil respectively which were significantly different from all experimental groups (p<0.001). The overall mean of length of leaves per plant was (15.98cm). As the effect of fertilizer, cultivars, and soil type was significant effect on leaf length increased progressively at all experimental cultivars, fertilizer and both soils even, if variations were observed among cultivars. This is since leaf length in forage grasses are greatly influenced by the developmental stage of the plant of the reproductive or vegetative stage. Leaf growth rate, increase prominently following flower induction and before any visible stem elongation (Crowder and Chheda, 1982).

Leaf length per plant was significantly affected by the interaction effect of fertilizer, type of cultivars and soil type in the current study was increased as level fertilizer increasing. The leaf length increased in the current finding at both soil types and all cultivars are in line with Biniyame Mihret et al. (2018) who reported that the length of leaves per plant of desho grass

55

was significantly affected by fertilizer type. The same author reported that the mean of leaf length was (30.48, 27.37 and 19.87cm) for NPS, manure and without fertilizer, respectively. The longest leaf length from the current study was recorded from B.mutica (22cm) at black soil and Brachiaria Marandu (20.27cm) at red soil, which was longer than desho grass and the current results of all experimental grass cultivars (15.98), was significantly shorter than desho grass reported by the same author. This difference might due to genetic variation, environmental conditions, altitudes, soil type, harvesting age, and management practices as well as interaction effects of cultivars, soil type, and fertilizer. Similarly the current result is in line with Nafatul et al.(2015) reported that as the level of NKP fertilizer increases from (0 to150 to 300 kg/ha), the mean leaf length of Brachiaria Marandu was slowly increased from (12.32 to 12.4 to 14.02 cm) respectively at 60 day harvesting age which was established by seed propagation system. The mean leaf length per plant (LLPP) of current interaction results from all cultivars (15.98cm) was longer than Brachiaria Marandu reported by the same author. This difference attributed to accesstion variation of grass species, environmental condition, harvesting age and management systems where both studies were conducted.

On the other hand, interaction effects of type of cultivars, soil type and fertilizer was affects the length of leaves per plant in the current study are in line with Brachiaria brizantha grass ecotypes as reported by Wubetie Adnew et al. (2018b). The overall mean of the longest length of leaves per was measured from Eth. 13809 (20.1 cm) followed by Eth. 13726 (17.29 cm) and Eth. 1377 (16.22 cm). While the overall mean of the longest leaf length per plant was recorded from midland (20.22 cm) followed by lowland (18.93 cm) and highland (14.47 cm). The current result of all experimental cultivars except Brachiaria Marandu and La Liberated at both soil types without fertilizer were significantly longer than as reported by the same author. This difference might come from genetic variation of grass species, environmental conditions and management practices where both studies were conducted as well as interaction effects of the current study. Generally, this current result showed that the type of cultivars, fertilizer level, and soil types had a significant effect on leaf length per plant in the current study.

56

4.1.5. Number of roots per plant

The interaction effects between fertilizer, soil type and type of cultivars on the number of roots per plant (NRPP) of Brachiaria grass in the current study were highly significant (p<0.001) (Table 4.1). The maximum number of roots per plant were obtained from hybrid Mulato-II (163.33) at red soil and Brachiaria Marandu (159.33) at black soil with fertilizer followed by hybrids of Mulato-II (154.36) and hybrid Mulato-I (154.76) with fertilizer at black soil. The intermediate number of roots per plant were recorded from Brachiaria La Liberated (149.5) at red soil, hybrid Mulato-I (147.3) at red soil and Brachiaria Marandu (148.7) at black soil with fertilizer while the minimum number of roots were observed from Mutica (111 and 105.5) without fertilizer at black and red soil respectively. The overall mean number of roots per plant were obtained from all experimental groups was (138.53).

Plant roots also keep the soil in place, reduce water leaching and soil erosion and key for soil phytoremediation. Root mass is heritable in perennial ryegrass at levels that will allow breeding for a larger number of root systems and progress is now being made towards changing the shape of root systems towards deeper rooting patterns (Crush et al., 2010). Generally, the number of roots per plant was increased, as interaction effects of fertilizer, soil type, and type of cultivars. Numbers of roots per plant were increased as 150 kg/ha, fertilizer was applied in the current study at both soil types and all cultivars. Root density in the soil profile is strongly dependent on the total number of roots per hill. Root spring out from node and the total number of nodes per plant determine the total number of roots even if the varietal differences in the average number of roots were observed (Akman and Topal, 2014). The roots, in turn, uptake water, and nutrients that the plants need to maintain turg or pressure and that assist with photosynthesis, as well as with promoting vegetative growth, flowering and fruiting (De Boer, 2016).

Fertilizer, type of cultivars and soil type had a significant interaction effect on the number of roots per plant in the current study are in line with Biniyame Mihret et al. (2018) who indicated that the fertilizer type significantly affects the number of roots desho grass in highlands of northwestern Ethiopia. The highest mean number of roots per plant were for NPS (116) followed by manure (94) and while the lowest was for without fertilizer (71). The maximum number of roots per plant in the current result were from Brachiaria hybrid Mulato-II (211) at red soil and the minimum Brachiaria hybrid of Mutica (137) at red soil,

57

of which was very higher than desho grass as reported by Biniyame Mihret et al. (2018) at 120 days of harvesting age. The overall mean number of roots in the current study were (161.14) higher than as reported by the same author. This difference might be the result of genetic variation, altitudes, soil type, level of fertilizer and management practices as well as interaction effects where the current study was conducted. The number of roots per plant (139.7) were counted from Brachiaria mutica grass at 90 days of harvesting age as reported by (Mimila Zemene, 2018). The overall mean number of roots counted from the current interaction result were (138.52) of all experimental cultivars of Brachiaria are lower and comparable with Brachiaria mutica grass as reported by the same author at the same 90 days of harvesting age. The number of roots per plant from B.mutica were (133.67 and 134.53) at black and red soil respectively in the current study was lower than the same cultivar B.mutica as reported by the same author. However, the number of roots per plant from Brachiaria Marandu, Brachiaria La Liberated, hybrid Mulato-II and Mulato-I results were higher than as reported by the same author at B. mutica grass. This difference might be coming from genetic variation experimental cultivars, environmental conditions, soil type, level of fertilizer, and management practices where the current experiment was conducted at soil type, fertilizer level, and type of cultivars interaction effect.

4.1.6. Length root per plant

The interaction effects between fertilizer, soil type and type of cultivars on length root per plant (LRPP) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.1). The longest length of root per plant was recorded from hybrid Mulato-II (15.6cm) at black soil and hybrid Mulato-II (14.07cm) at red soil with fertilizer, they were significantly different each other followed by La liberated (14cm), Marandu (13.8cm) and from mutica (13.6cm) with fertilizer at black soil and were not significantly different each other. The intermediate length of the root was recorded from La liberated (12.6cm), Marandu (11.9cm), Mutica (11.8cm) at red soil and hybrid Mulato-I (11.9cm) at black soil with fertilizer. While the shortest length of the root was obtained from cultivars of Mutica (5.87cm) at red soil without fertilizer, La liberated (7.6cm) and Marandu (7.8cm) at both soil types without fertilizer and they were not significantly different each other (p<0.001). The overall mean of length of roots per plant measured from all experimental groups was (10.69cm).

58

Fertilizer, type cultivars, and soil type had significant interaction effect on the length of roots per plant in the current study are in agreement with desho grass as reported by Biniyame Mihret et al.(2018) indicated that the fertilizer type and level of fertilizer significantly affected the length of roots per plant desho grass in highlands of northwestern Ethiopia. The longest mean length of roots per plant was recorded for NPS (26.05cm) followed by manure (25.36cm) and while the shortest was for without fertilizer (22.84cm). The longest length of roots per plant in the current study was from hybrid Mulato-II (15cm) and hybrid Mulato-II (14.07) of which is shorter than desho grass as reported by the same author at 120 days of harvesting age. The overall mean length of roots per plant in the current result was (10.69cm) significantly shorter than as reported by the same author at the same 120 days of harvesting stage. This difference might be coming from genetic variation, environmental condition, harvesting age, fertilizer application and management systems where the current experiment was practiced. Generally, it indicated that the length of the roots increased with increasing level of fertilizer from zero (0) to 150 kg/ha in the current study at both soils and all cultivars. An increase in root length increase the absorbing surface, thus they may be important for water and mineral absorption. The explanation was that the uptake of a different nutrient in competing plants was proportional to the root length, as the greater the root length, the shorter the distance the nutrient has to travel to the root (Crush et al., 2010). In Mediterranean environments, greater root length and deeper roots are more important for higher water uptake (Akman and Topal, 2014).

4.1.7. Leaf to stem ratio

The interaction effects between fertilizer, soil type and type of cultivars on the leaf to stem ratio of Brachiaria grass cultivars in the present study was highly significant (p<0.001) (Table 4.1). The highest leaf to stem ratio was measured from Marandu (2.27) at red soil with fertilizer and hybrid Mulato-II (2.17) at black soil with fertilizer followed by hybrid Mulato-II (1.98) and hybrid Mulato-I (1.96) with fertilizer at red and black soil, respectively. The intermediate leaf to stem ratio was observed from Marandu (1.84) and hybrid Mulato- I (1.82) with fertilizer at black and red soil respectively. The lowest leaf stem ratio was optioned from B.mutica cultivar (1.04 to 1.13) with and without fertilizer at both soil types. The overall mean of a leaf to stem ratio from all experimental groups was (1.63).

59

Leaf to stem ratio in the current study was significant interaction among cultivars, soil types, and fertilizer. The highest leaf to stem ratio was measured from Marandu and hybrid Mulato-II and while lowest was optioned from B.mutica grass at both soil types. This significant difference among treatment groups attributed to the genetic potential of cultivars' ability to respond at the same management system and environmental conditions and production of more number of leaves per plant. Mimila Zemene (2018) reported that B.mutica grass had a (1.37) leaf to stem ratio at 90 days of harvesting age. The overall mean of a leaf to stem ratio from current all experimental groups (1.63) was significantly higher than as reported by the same author at the same harvesting age. This difference might come from the genetic variation of grasses, environmental condition, altitudes, soil type, level of fertilizer, rainfall, temperature and management practices, however, with the same B.mutica cultivar leaf to stem ratio of the current result was nearly comparable with finding of the same author at same harvesting ages.

Bimrew Asmare et al. (2017) reported that desho grass had a (1.24) leaf-to-stem ratio at 90 days of harvesting age. The current interaction result of the leaf to stem ratio was higher than desho grass as reported same author while this result was comparable with the current result of B. mutica cultivar. This difference might be coming from a genetic variation of grasses, altitudes, soil type, level of fertilizer, rainfall, temperature and management practices. Generally, fertilizer, type of cultivars and soil type had an interaction effect on the leaf to stem ratio of experimental Brachiaria grass cultivars, even though there were inconsistent results observed. Higher leaf to stem ratio was recorded from the interaction effect of cultivars, soil type and fertilizer level at all experimental cultivars as well as soil types. It is an important factor affecting diet selection, nutritive value and intake of forage (Smart et al., 2004).

4.1.8. Dry matter yield ton per hectare

The interaction effects between fertilizer, soil type and type of cultivars on dry matter yield (DMY) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.1). The highest dry matter yield was recorded from B.mutica (20.37t/ha) at black soil with fertilizer and it was significantly higher than all other treatment groups followed by the hybrid Mulato-II (18.62t/ha) at back soil with fertilizer and B.mutica (17.2t/ha) at red soil with fertilizer. The intermediate dry matter yield was recorded from hybrid Mulato-II

60

(13.29t/ha) at red soil, hybrid Mulato-I (11.55 and 11.07t/ha) with fertilizer at red and black soil respectively. The lowest dry matter yield was recorded from La Liberated (2.75 and 3.47t/ha) without fertilizer at black and red soil respectively. The overall mean of dry matter yield was (8.45t/ha). Generally the highest yield from all experimental cultivars of Brachiaria grasses at fertilizer treatment groups at both soil types could be attributed to the formation of additional tillers developed which conveyed an increase in leaf formation, leaf elongation and stem development as interaction effect of fertilizer, cultivars and soil type (Crowder and Chheda, 1982).

Fertilizer, cultivars and soil type had a significant effect on the total dry matter yield of Brachiaria grass in the current study is in line with Biniyame Mihret et al. (2018) reported that DMY of desho grass was significantly affected by fertilizer type and level fertilizer. The highest mean DMY ton per hectare was from NPS (24.08) followed by manure (14.05), while the lowest DMY was recorded from control or without fertilizer (8.55). The highest DMY in the current study was recorded from B.mutica (20.37) and hybrid Mulato-II (18.62) which was significantly higher than desho grass of manure and control groups as reported by the same author. However, the current result from these cultivars and the overall mean was lower than desho grass of NPS treatment group as reported by the same author. This difference might be due to the genetic variation of grass, environmental conditions and management system like harvesting age where the current experiment was conducted. Similarly results with Worku et al. (2017) indicated that there was significant difference (p<0.05) in DMY of desho grass ton per hectare was increased as the fertilizer rate increase from 100 kg /ha (8.32t/ha) to 150 kg/ha (9.09t/ha) to 200kg/ha (12.73t/ha) to 250kg/ha (13.6t/ha). The highest result from cultivars of (mutica, hybrid Mulato-II, and hybrid Mulato-I) was higher than desho grass as reported by the same author. This difference might be coming from a genetic variation of grasses, environmental conditions, level of fertilizer, interaction effect, sampling procedures etc. Nafatul et al. (2015) reported that as the levels of NPK fertilizer increases from 0 to 150 to 300 kg/ha the DMY of B.brizantha Marandu grass was increased linearly from 0.78 to 1.13 to 0.99 t/ha respective of fertilizer at 60 days of harvesting age. The overall mean results from all experimental cultivars of the current study were very higher than B.brizantha Marandu as reported by the same author. This higher difference might be coming from the environmental condition, genetic variation,

61

harvesting age difference, establishment method and type of fertilizer in addition to interaction effects where the current experiment was conducted.

On the other hand dry matter yield was affected by interaction effects of type of cultivars or species, fertilizer and soil type in the current study is in line with Mustering et al. (2014) reported that DMY kg per hectare was significantly affected by types of species and cultivars of Brachiaria grasses. The highest DMY was recorded from Marandu Mulato-I (0.79kg/m2) followed by mutica (0.55kg/m2) while the lowest DMY was recorded from B. brizantha (0.51kg/m2) at 60 days of harvesting after establishment phase. Similarly, Mutimura and Everson (2012) noted that types of species or cultivars affected DMY. The highest DMY ton per hectare was from hybrid Mulato-II (5.13) and Mulato-I (5.03) followed by B. decumbens (4.79) and B.brizantha Marandu (4.58). The overall mean results from all experimental cultivars with fertilizer except Brachiaria La Liberated in the current study were significantly higher than as reports of the same both authors. This difference might be a difference in the environmental conditions, management systems, genetic variations, where both studies were conducted. The overall results from all current experimental cultivars except Brachiaria La Liberated without fertilizer was very higher than Brachiaria brizantha ecotypes as reported by Wubetie Adnew et al.( 2018b) at 90 days of harvesting ages which were 3.63 t/ha. Moreover, Mimilia Zemene (2018) indicated that DMY ton per hectare of Brachiaria mutica grass was significantly affected by harvesting age. The highest (16.17) DMY ton per hectare was recorded from 120 days of harvesting followed by 90 (14.79), while the lowest DMY was observed at 60-day harvesting age (5.08 t/ha). The mean of DMY resulted from B. mutica in the current study was higher than the same species of B. mutica grass as reported by the same author at the same 90 days of harvesting age. This difference might be coming from soil type and fertility, type of fertilizer and level of fertilizer, management systems. The highest herbage yield from the interaction effect of fertilizer, cultivars and soil type Brachiaria grass in the current study was recorded. This might become from the rapid increase in the tissues of the plant, development of additional tillers and leaf formation, leaf elongation and stem development with increasing fertilizer level from 0 to 150 kg/ha in the current study as reported by Biniyame Mihret et al.(2018).

62

Table 4. 1 Effect of fertilizer, cultivars and soil type on morphological characteristics and dry matter yield of Brachiaria grasses

Variables Parameters ST FER Cultivars PH (m) NTPP NLPP LLPP NRPP LRPP LSR DMY (count) (count) (cm) (count) (cm) T/HA Mutica 1.55a 40bc 333.3b 22a 133.67ef 13.6bc 1.04hi 20.37a Mulato-II 0.82d 46a 380.93a 20.2b 154.3bc 15.6a 2.17ab 18.61b With Mulato-I 0.52fg 36.3efg 233.7de 17.9cd 154.7bc 11.9de 1.96abc 11.07e Marandu 0.51fg 29.8j 185.03fg 15.9e 148.7cd 13.8bc 1.84bcd 7.13g Black La Liberated 0.49fg 26.2k 153.49ghij 13.6hi 138.3e 14.07b 1.76bcde 5.7jkl Out Mutica 1.12c 32.2hi 173.03gh 18c 113h 10.13fg 1.14 ghi 6.95gh Mulato-II 0.55f 36.3efg 185.2fg 17.5cd 125.3g 11.13ef 1.52 defg 6.5hi Mulato-I 0.37ij 31.03ij 157.27ghi 15.2ef 138e 8.53hi 1.47efgh 5.33kl Marandu 0.32jk 26.67k 143.15hij 14.9fg 137.3e 7.8i 1.78bcde 3.85n La Liberated 0.35jk 24.9k 124.07j 11.17k 132.67ef 7.6i 1.3fghi 2.75o With mutica 1.45b 41.67b 264.0c 17.03d 134.53ef 11.8de 1.09i 17.19c Mulato-II 0.62e 45.27a 366.68a 13.07ij 163.33a 14b 1.98abc 13.29d Red Mulato-I 0.52fg 39cd 244.23cd 12.57j 147.3d 9.8fgh 1.82bcd 11.55e Marandu 0.5fg 37.67de 217.8de 20.27b 159.33ab 11.9de 2.27a 6.02ij La Liberated 0.46gh 36.03efg 154.49ghij 20.58b 149.5cd 12.6cd 1.77bcde 5.87jk Out Mutica 0.85d 20.33l 135.67ij 14.7fg 105.5i 5.87j 1.02i 5.27l

65

Table 4.1 (continued) Cultivars PH (m) NTPP NLPP LLPP NRPP LRPP LSR DMY (count) (count) (cm) (count) (cm) T/HA Mulato-II 0.42hi 34.8fg 163.52ghi 10.74k 138e 9.7gh 1.7cdef 8.2f Mulato-I 0.36ij 31.17ij 208.97ef 13.03ij 133.3ef 8.5hi 1.68cdef 5.65jkl Marundu 0.38ij 37.0def 222.07de 17.13cd 134 ef 7.8i 1.769bcde 4.44 m La Liberated 0.3k 34.0gh 208.5ef 14.13gh 129.67 fg 7.6i 1.39efghi 3.47 n Overall mean 0.62 34.32 212.76 15.98 138.53 10.69 1.63 8.46 SEM 0.04 0.86 9.19 0.41 3.42 0.34 0.05 0.63 CV 5.6 3.74 8.19 3.17 2.79 7.15 13.23 3.78 R2 0.993 0.974 0.963 0.983 0.952 0.948 0.805 0.997 SL *** *** *** *** *** *** *** ***

Means in the same column followed by different letters differ significantly at P 0.001. SL=Level of significance, ***P 0.001, NS=Non-significant, PH=plant height, NTPP= Number of tiller per plant NLPP=Number of leaves per plant, LLPP= Length leaf per plant, NRPP= Number of roots per plant, LRPP= length of root per plant, LSR=leaf to stem ratio, DMYT/HA=Dry matter yield ton per hectare, SEM= Standard error of mean.CV= Coefficient of variation, R2=Coefficient of determination, FER= Fertilizer

66

4.2. Effects of fertilizer, cultivars and soil type on the quality of Brachiaria grasses

The overall analysis result showed that there were significant interactions effects among fertilizer, type of cultivars and soil type. It is worthless to interpret the main effect of the current study. The interaction effects of fertilizer, cultivars and soil type of Brachiaria grasses on DM, ASH, OM, NDF, ADF, ADL, IVDMD, IVOMD, CP, CPY and ME were presented in (Table 4.2, 4.3 and Appendix 7.1.4). All quality parameters were significantly affected by the interaction effect of fertilizer, type of cultivars and soil type in the current study in (Table 4.2, 4.3). Generally, the overall nutritive value of a given forage is affected by soil nutrients (Tessema Zewdu et al., 2011), climate (Arzani et al., 2008) and associated factors. Besides, the inherent type of species and soil characteristics are important factors in altering the nature of forage nutritive value (Bimrew Asmare et al., 2017).

4.2.1. Dry matter content

The interaction effects between fertilizer, soil type and type of cultivars on dry matter content (DM %) of Brachiaria cultivars in the current study were a significant (p<0.001) (Table 4.2). The highest dry matter content was recorded from hybrid Mulato-II (94.04) at red soil without fertilizer followed by the hybrid Mulato-I (93.9) at both soil type without fertilizer and B. mutica (93.82 and 93.9) at red soil without fertilizer and at black soil with fertilizer respectively. The intermediate dry matter content was recorded from B.mutica (93.78) at black soil without fertilizer and red soil with fertilizer followed by B.brizantha Marandu (93.74) at red soil without fertilizer. While the lowest dry matter, the content was observed from hybrid Mulato-II (92.97) at black soil and Marandu (93.065) at red soil without fertilizer. The overall mean of dry matter content was (93.55).

Application of fertilizer, soil type and type of cultivars had significant (p<0.001) interaction effect on dry matter content of Brachiaria grasses in the current study is disagree with desho grass as reported by (Biniyame Mihret et al., 2018) on type of fertilizer had no significant effect on dry matter content at 120 days of harvesting age. Similarly, the current study is disagreed with (Abdi Hassan et al., 2015) reported that on Cenchrus ciliaris and Panicum maximum grasses application of nitrogen fertilizer had no significant effect on dry matter content under irrigation condition. This difference might be due to the type of fertilizer, the genetics of grasses, season and management system where studies conducted. Additionally,

67

types of cultivars or species, fertilizer and soil type interaction effects on dry matter content in the current study is in line with Wubetie Adnew et al. (2018b) reported that dry matter content was significantly affected by the type of ecotypes Brachiaria brizantha grass. The highest dry matter content was recorded from Eth. 13809 (37.75%) followed by Eth.13727 (36.71%) and Eth.1377 (35.71%). The same author reported that agro ecology had a significant effect on the dry matter content of Brachiaria brizantha ecotypes. The highest dry matter content was reported at lowland (37.77%) followed by midland (36.61%) and highland (35.79%). The overall mean results from all cultivars in the current study were higher than Brachiaria brizantha ecotypes as reported by Wubetie Adnew et al. (2018b). The overall mean of dry matter content in the current result was (93.55%) lower than B .mutica grass as reported by Mimilia Zemene (2018). The overall mean of dry matter content in the current interaction result study was (93.55%) higher than the same species B.mutica grass as reported by the same author at the same 90 days of harvesting age. This difference might be coming from soil condition, genetics, fertilizer, interaction effect and drying methods of samples of authors where both studies were conducted.

4.2.2. Ash and Organic matter

The interaction effects among fertilizer, soil type and type of cultivars on ash (%) and organic matter content (OM %) of Brachiaria grass in the current study were highly significant (p<0.001) (Table 4.2). The highest ash content was recorded from La Liberated (13.58) at red soil with and without fertilizer and from B. brizantha Marandu (13.37) at red soil without fertilizer. The intermediate ash content was recorded from La Liberated (12.98 and 12.92) at black soil with and without fertilizer. The lowest ash content was from B.mutica (8.53) at red soil with and without fertilizer respectively and at black soil without fertilizer. While the highest organic matter content was at the lowest ash content and the lowest organic matter content was at the highest ash content. The overall mean of ash and organic matter content were (11.61 and 88.39) percent respectively.

The highest ash content in the some Brachiaria grass cultivars in the current study; might be an indication of the high mineral concentration of grasses as a result of interaction effects among fertilizer, type of cultivars and soil type. The overall mean interaction results on the mineral (Ash) content of the grass was lower with at fertilizer, soil type, and cultivars interaction effect results. This might be the mineral content declines due to a natural dilution

68

process and translocation of minerals to the roots (Minson, 1990). The concentration of minerals in forage grasses varies from spices to species; this could be factors like, morphological fractions, climatic conditions, soil characteristics and fertilizer levels (McDowell and Valle, 2000).

Fertilizer application, type of cultivars and soil type had a significant interaction effect on the ash and organic matter content in the current study. The current study is in agreement with Biniyame Mihret et al, (2018) reported that type of fertilizer and level of fertilizer had a significant effect on the ash and organic matter content of desho grass. The highest ash content was recorded at manure (17.77) followed by without fertilizer (16.41) and NPS fertilizer was the lowest ash content (15.76) while the highest organic matter content was recorded at NPS followed by manure and the lowest was at without fertilizer. The current results of ash content from all experimental cultivars were lower than desho grass as reported by the same author at 120 days of harvesting age. This discrepancy due to come from the genetic variations of grasses, soil type, altitude and management systems (Campos et al., 2013). Similarly, the current result is in agreement with Abdi Hassan et al. (2015) reported on Cenchrus ciliaris and Panicum maximum grasses of ash content and organic matter. It was significantly affected by nitrogen fertilizer levels, the level of fertilizer increases from 0 to 50 to 100 kg/ha as the content was decreased from (14.43 to 11.85 to 11.89) respectively, while the organic matter content was increased as the level of fertilizer increased. The lowest total ash content of forages from fertilized Brachiaria grass cultivars in the current result, which brings about earlier dilution and translocation of different minerals associated with a vegetative portion of the plant (leaf portion) to roots at fertilizer as described by Maynard et al. (1981).

Type of cultivars or species, fertilizer, and soil type had a significant interaction effect on the ash and organic matter content of Brachiaria grass in the current study is disagreement with Brachiaria brizantha grass ecotypes as reported by Wubetie Adnew et al. (2018b). There was no significant difference among the three Brachiaria brizantha ecotypes of Eth. 13726, Eth.13809 and Eth.1377. However, agro ecology had a significant effect on ash and organic matter content of Brachiaria brizantha ecotypes as reported by the same author. The highest ash and lowest organic matter content was recorded from lowland (12.22 and 87.88) followed by highland (12.17 and 87.83) and midland (10.09 and 89.91) respectively. The current result from cultivar La Liberated at both soil without fertilizer, Brachiaria 69

Marandu without fertilizer was significantly higher than Brachiaria brizantha ecotypes as reported by the same author and lower than Brachiaria brizantha ecotypes the most other cultivars with fertilizer at both soil types. This difference could become from variation of accsstions, environmental conditions and management practices differences as well as their interaction effects. Mineral (ash) nutrients in the feed play major roles in the body function of the overall animal production and productivity activity including skeletal development and maintenance, energy, milk production and body function (Rasby et al., 2011). Concentrations of mineral nutrients in plants were affected by the environment, management applied to the forages (El-Nashaar et al., 2009).

4.2.3. Neutral detergent fiber

The interaction effects between fertilizer, soil type and type of cultivars on neutral detergent fiber content (NDF %) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.2). The highest NDF content was from B.mutica (78.97 and 77.9) at red soil with and without fertilizer and from B.mutica (78.98) at black soil without fertilizer. The intermediate NDF content was recorded from La liberated (76.22) at black soil without fertilizer and from B.mutica (73.93) at black soil without fertilizer. While the lowest NDF content recorded was from hybrid Mulato-II (65.75) at black soil with fertilizer, hybrid Mulato-I (65.07) at both soils without fertilizer and Marandu (65.34) at red soil without fertilizer. The overall mean of NDF content was (70.134).

The highest overall mean of NDF content result was recorded at both soil types in most experimental Brachiaria grass cultivar groups in the current study without fertilizer groups. This indicates that as the level of fertilizer increases from zero (0) to 150 kg/ha, the NDF content decreases. This may elucidate that the fertilizer application effect expands the plant growth, raise new leaves, tillers and shoots, which minimizes the NDF content in the fertilizer level increased, but there is no rejuvenation of leaves and tillers in the non-fertilizer treatments as a result plant tissue matures and accumulate more NDF. This attributed to decrease NDF content, increased growth rate of new leaves and shoot which are lower in plant structural components because of urea fertilizer (Abdi Hassan et al., 2015).

Fertilizer, soil type, and cultivars had a significant interaction effect on the NDF content Brachiaria grass in the current study. This current interaction result is in agreement with

70

desho grass as reported by Biniyame Mihret et al. (2018) that the NDF content of desho grass was affected significantly by fertilizer type and level of fertilizer. The highest NDF content was observed without fertilizer (61) followed by manure application (55) while the lowest observed at NPS fertilizer (51). The overall NDF content from the current result was (70.13) significantly higher than desho grass as reported by the same author at 120 days of harvesting age. This higher difference might be coming from genetics, altitude, soil type, harvesting age and environmental conditions besides interaction effects. Similarly NDF content form all experimental cultivars current result was comparable with Cenchrus ciliaris and Panicum maximum grasses were decreased as the nitrogen level of fertilizer increase from 0 (70.89) to 50 (68.52) to 100 kg/ha (61.03) as reported by Abdi Hassan et al., 2015) under irrigation condition.

In the current study, NDF content significantly affected by the type of species, fertilizer, and soil type interaction effect of Brachiaria grass agrees with Mustaring et al. (2014) reported that NDF content of Brachiaria grass was affected by type and species of grass. The highest NDF was observed form B.mutica grass (71.96%) followed by Brachiaria brizantha (65.28) while the lowest from Brachiaria hybrid Mulato- I (63.66) at similar harvesting age (60 days) in lands of central Sulawesi, Indonesia. In the current interaction result B.mutica (73.93 to 78.97) and Brachiaria hybrid Mulato-I (65.06 to 67.28) at 90 days of harvesting age of which significantly higher than same cultivars as reported by the same author at 60 days of harvesting age. This difference might be coming from environmental conditions, soil type, and fertility, harvesting age, fertilizer where both studies were conducted. According to Van Saun (2006), forage grass with less than 50% NDF described as high nutritive value whereas NDF greater than 60 % considered as low nutritive value. Therefore the current interaction result of NDF content form all experimental cultivars classified as low nutritive value forage according to the same author classification.

4.2.4. Acid detergent fiber

The interaction effects among fertilizer, soil type and type of cultivars on acid detergent fiber content ADF %) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.2). Significantly, the highest ADF content was from B.mutica (63.66) at black soil and red soil without and with fertilizer respectively followed by the same cultivar (58.51) at red soil without fertilizer. The intermediate ADF content was from B.mutica

71

(56.35) at black soil without fertilizer and from hybrid Mulato-II (55.48) at black soil without fertilizer. While the lowest ADF content was from B.brizantha Marandu (45.59) at red soil without fertilizer. The overall mean of ADF content was (52.48).

Generally, as fertilizer, soil type and cultivars interaction effect on the ADF content was lower at all treatment groups. The fertilizer application level increases from zero to 150 kg/ha, the ADF content decreased in the current result of the interaction effect among soil type and cultivars. This due to associated with the fertilizer application promotes the growth of new leaves and shoots resulting in low lignin, which compensates for the increase in lignin content of other tissues. When lignin is lowered it has always produced a marked increase in the digestibility of the plants and lignin are highly resistant to chemical and enzymatic degradation and are not appreciably broken down by the micro flora in the ruminant digestive tract (Ranjhan,1993). Acid detergent fiber is the percentage of highly indigestible and slowly digestible material in a feed or forage. Higher forage ADF results in reduced digestibility dry matter because of increased lignification of cellulose in the latter stage of the plants (Depeters, 1993).

Fertilizer, soil type, and type of cultivars had a significant interaction effect on the ADF content Brachiaria grass in the current study. The current finding is in agreement with (Biniyame Mihret et al., 2018) reported that ADF content desho grass was significantly affected by fertilizer type and level fertilizer. The highest ADF was observed without fertilizer (52) followed by manure application (44) while the lowest observed at NPS fertilizer (41).The overall mean of ADF (52.48) from the current study significantly higher than desho grass with NPS and manure treatment groups as reported by the same author at 120 days of harvesting stage. Similarly, the current result was in line with ADF content Cenchrus ciliaris and Panicum maximum grasses. ADF decreased as the nitrogen level of fertilizer increase from zero (50.37) to 50 (44.91) to 100 kg/ha (42.89) (Abdi Hassan et al., 2015) under irrigation conditions. This was lower than the current result, might be due to associated with the genetic variation of grasses, altitude differences, soil type and fertility, harvesting age and fertilizer type where the previous and the current experiments were conducted.

ADF content of Brachiaria grass was affected by type and species of grass, fertilizer and soil type interacted effect in the current study is in line with Mustaring et al. (2014). The

72

highest ADF was from B .mutica grass (46.09) followed by hybrid Mulato-I (38.79) while the lowest was from Brachiaria brizantha (38.21) at similar harvesting age 60 days in the central lands of Sulawesi, Indonesia. The current result of ADF content form all experimental cultivars was higher than as reported by the same author at the same species of Brachiaria grass cultivars. This discrepancy of the current finding from previous reports might be coming from the environmental conditions, soil type and fertility, harvesting age, and fertilizer where the current study was conducted. The current result of ADF content was lower than B.mutica, B.brizantha ecotypes and desho grass as indicated by previous reports (Mimilia Zemene, 2018; Wubetie Adnew et al., 2018b; Genet Tilahun et al., 2017). This difference due to genetics, fertilizer and environmental conditions. The structural cell wall components of grass were increased as the plant gets matured, this might be related to photosynthesis components are converted to structural components at the expense of soluble carbohydrates (Ammar et al., 2010).

4.2.5. Acid detergent lignin

The interaction effects between fertilizer, soil type and type of cultivars on acid detergent lignin content (ADL %) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.2). The highest ADL content was from B.mutica (12.09) at both soils without fertilizer followed by the same cultivar (10.8) at black soil with fertilizer and from hybrid Mulato-I (10.77) at red soil without fertilizer. The intermediate ADL content was from Marandu (10.23 and 10.31) at black soil with and without fertilizer respectively, hybrid Mulato-I (10.7) at black soil with fertilizer, from hybrid Mulato-II (10.43) at black soil without fertilizer and Mutica (10.24) at red soil without fertilizer. While the lowest ADL content was recorded from hybrid Mulato-II (9.14) at both soils without fertilizer and La Liberated (8.94) at black soil without fertilizer. The overall interaction mean of ADL content was (10.04).

The application of fertilizer, soil type, and cultivars had a significant interaction effect on the ADL content of Brachiaria grasses in the current study at both soil types. This current study is in line with desho grass as reported by (Biniyame Mihret et al., 2018) indicated that ADL content of desho grass was significantly affected by fertilizer type and level of fertilizer. The highest ADL content was without fertilizer (13.5) followed by manure application (9.8) while the lowest observed at NPS fertilizer (8.3) at 120 days of harvesting

73

age. The overall means of ADL content (12.09) which was higher and lower than desho grass as reported by the same author with and without fertilizer respectively in the current study. Similarly ADL content of current result is in line with Cenchrus ciliaris and Panicum maximum grasses were decreased as the nitrogen level of fertilizer increase from 0 (10.26) to 50 (9.3) 100 kg/ha (6.59) (Abdi Hassan et al., 2015) under irrigation condition. The NDF content form all cultivars of current interaction result was higher than as a reported same author. This attributed to the genetic variation of grasses, harvesting age, growing season, interaction effect and environmental conditions where both studies were conducted. Generally, as fertilizer applied on experimental cultivars the ADL content was lower at both soil types due to interaction effect. This is due to the formation of fresh and new more leaves than without fertilizer (Abdi Hassan et al., 2015).

The ADL content Brachiaria grass was affected by the interaction of type of cultivars or species, fertilizer and soil type in the current study is in line with Mustaring et al. (2014) reported that ADL content of Brachiaria grass was affected by type and species of grass. The highest ADL was observed at B.mutica (11.58) followed by Brachiaria brizantha (8.48) while the lowest was recorded from hybrid Mulato-I (8.18) at similar harvesting age (60 days) in lands of central Sulawesi, Indonesia. The current results at the same cultivars of Mutica and hybrid Mulato-I was higher than as reported by the same author at 60 days of harvesting age. This difference might be coming from the genetics of grass species, environmental conditions, altitude, soil type and fertility and harvesting age where both studies were conducted as well as the contribution of the interaction effect. Forages with higher ADL content had low overall digestibility of the feed by limiting nutrient availability (Van Soest, 1994). This high lignin content of experimental cultivars might be the ability to bind the cellulose and hemicellulose and prevent them from digested and utilized efficiently by the rumen microbes. Therefore, forages with lower ADL concentrations are more desirable for livestock production and productivity (Ansah et al., 2010). Mimilia Zemene (2018) and Wubetie Adnew et al. (2018b) reported that ADL content of B.mutica and Brachiaria brizantha grass ecotypes were significantly affected by harvesting stage respectively. ADL increases with the advancing maturity of forages from 60 to 90 to 120 harvesting ages as reports of the same authors.

74

Table 4. 2 Effects of fertilizer, soil type and different cultivars on dry matter content, ash, organic matter and fiber contents of Brachiaria grass cultivars

Variables Parameters (Dry matter basis) ST FER Cultivars DM (%) ASH (%) OM (%) NDF (%) ADF (%) ADL (%) WF mutica 93.89ab 10.13f 89.872b 73.93c 56.35c 10.78b Mulato-II 93.37fgh 12.71cd 87.29ed 65.75i 48.58g 9.64de

Mulato-I 93.61cdef 11.52e 88.48c 67.28fghi 49.13fg 10.67bc Marandu 93.29ghi 12.76cd 87.24ed 71.3d 53.53d 10.23c Black La Liberated 93.4fgh 12.98c 87.024e 66.51ghi 52.46de 9.48de WO mutica 93.78bcd 8.56g 91.44a 78.97a 63.66a 12.09a Mulato-II 92.97j 9.8f 90.2b 68.98def 55.68c 10.43bc

Mulato-I 93.9ab 11.35e 88.65c 65.07i 49.21fg 9.14ef Marandu 93.41fgh 12.3d 87.7d 70.89d 53.69d 10.31bc

La Liberated 93.53efg 12.92c 87.08e 76.22b 50.28fg 8.94f Red WF mutica 93.78bcd 8.56g 91.44a 78.97a 63.66a 12.09a Mulato-II 93.28hi 11.391e 88.61c 68.46efg 49.78fg 9.78d Mulato-I 93.55def 11.18e 88.82c 66.09hi 49.39fg 10.77b Marandu 93.07ij 12.73cd 87.27ed 68.08efgh 49.19fg 9.43def La Liberated 93.4fgh 13.58b 86.42f 69.05def 50.9ef 9.34def WO mutica 93.82abc 8.56g 90.41b 77.91ab 58.51b 10.24c

75

Table 4.2 (continued) Cultivars DM (%) ASH (%) OM (%) NDF (%) ADF (%) ADL (%) Mulato-II 94.04a 10.92e 89.079c 69.74de 49.84fg 9.53de Mulato-1 93.9ab 11.35e 88.65c 65.07i 49.21fg 9.14ef Marandu 93.75bcde 14.37a 85.63g 65.36i 45.591h 9.35def La Liberated 93.39fgh 13.58b 86.42f 69.05def 50.9ef 9.34def Overall Mean 93.56 11.6 88.39 70.13 52.48 10.04 SEM 0.05 0.26 0.26 0.74 0.77 0.14 CV 0.11 2.31 0.3 1.4 1.5 2.1 R2 0.937 0.986 0.986 0.976 0.986 0.972 Sig *** *** *** *** *** ***

ST= soil type, FER= Fertilizer, DM= Dry matter content, ASH= ash, OM= organic matter, NDF=Neutral detergent fiber, ADF=Acid detergent fiber, ADL=Acid detergent lignin, R2 = Coefficient of determination, CV= Coefficient of variation, Sig = Significance level, WF= with fertilizer, WO =without fertilizer

76

4.2.6. Crude protein

The interaction effects between fertilizer, soil type and type of cultivars on crude protein content (CP %) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.3). The highest CP content was from Marandu (14.02 and 13.78) with fertilizer at red and black soil respectively followed by hybrid Mulato-II (13.21 and 13.04) at red and black soil with fertilizer. The intermediate CP content was from hybrid Mulato-I (12.7 and 12.6) at black and red soil with fertilizer respectively followed by La Liberated (11.6 and 11.31) at black and red soil respectively. The lowest CP content was from B.mutica (7.25 and 7.2) without fertilizer at red and black soil respectively. The overall mean of CP content was (10.5).

The CP content is one of the most important criteria to determine the nutritional quality of livestock feeds; this is due to the level of CP increases, the dry matter intake by livestock and rumen microbial growth would also increase (Chanthakhoun et al., 2012). The difference in CP content of Brachiaria grass in the current study can explained by the inherent characteristics of each of Brachiaria species associated with the ability to extract and accumulate nutrients from the soil and fix nitrogen from the atmosphere as reported by (Yusuf and Muritala, 2013). According to Van Soest et al. (1991), CP level of 7.5 % is required for rumen function. The CP content recorded from all cultivars except Brachiaria Mutica at both soil types without fertilizer was satisfied above minimum requirement of ruminant animals. The CP content at both soil types from all cultivars is satisfying above minimum requirements of ruminant livestock production with fertilizer experiment groups.

The application of fertilizer, type of cultivars and soil type had a significant interaction effect on the CP content Brachiaria grass in the current study. As fertilizer applied, the CP content was higher at all experimental cultivars in both soil types. The levels of fertilizer increased from zero (0) to 150 kg/ha in the current study, the CP content also increased in all experimental cultivars at both soil types. This might be because of continued application fertilizer, allowed to continuous sprouting of the grasses of new leaves, which was a bit fresh even during the harvest of forage biomass (Abdi Hassan, et al., 2015). The current result is in line with desho grass as reported by (Biniyame Mihret et al., 2018) the CP content of desho grass was significantly affected by fertilizer type and level of fertilizer. The highest CP content was from NPS fertilizer (10.95) followed by manure application (10.04) while

77

the lowest observed at without fertilizer (9). The overall mean of CP content from the current interaction result from all cultivars was higher than desho grass as reported by Biniyame Mihret et al. (2018) except B.mutica. This difference might become from genetics, harvesting age, environmental conditions, soil type, and fertility interaction effects of soil, fertilizer, and cultivars. Similarly, Abdi Hassan et al. (2015) reported that CP content of Cenchrus ciliaris and Panicum maximum grasses were increased as the nitrogen level of fertilizer increases from 0 kg/ha ( 13.54) to 50 kg/ha (17.34) to 100 kg/ha (19.84) under irrigation condition. The overall current result of CP from all experimental cultivars was lower than above grass species as reported by the same author. This difference due to genetic variations of grass species, type of fertilizer and production season, the interaction effect of which is his experiment was done in the dry season at irrigation condition.

The current result of CP content was higher than as reported by the same author within the same species Brachiaria cultivars of the current study. Similarly, Mutimura et al. (2017) reported that the type of species or cultivars of Brachiaria grass in line with the current study affected CP content. The highest CP content was recorded from Brachiaria Decumbens (16.7) and the lowest was from Brachiaria hybrid Mulato- I (13.8) at 90 days of harvesting age. The current result from all cultivars was lower CP content than as reported by the same author. Susan et al. (2015) reported also the type of species and cultivars had a significant effect on the CP content of Brachiaria grasses. The highest CP was recorded from Brachiaria Marandu (9.2) followed by Brachiaria La Liberated (7) and Brachiaria decumbens (4.9) at 60 days of harvesting age. The current result was higher than as reported by the same authors at the same species of Brachiaria cultivars. This difference from both authors might become from environmental conditions, management systems and application of fertilizer where both studies were conducted.

On the other hand, Wubetie Adnew et al. (2018b) reported that the CP content was different among the type of Brachiaria brizantha ecotypes. The highest CP content was recorded from Eth.13726 (12.36) followed by Eth.1377 (11.17) and the lowest was from Eth.13809 at 90 days of harvesting age. The current result of CP content from Marandu, hybrid Mulato- II, and hybrid Mulato-I was higher than as reported by the same author. This difference might become from the genetic variation of grasses, environmental conditions etc. Mimilia Zemene (2018) reported that harvesting age significantly affects the CP content of B.mutica. The CP content was increasing trend (6.16 to 9.46 to 13.52) as harvesting age decreased 78

form (120 to 90 to 60) days respectively. The current result is higher than the same species of B.mutica as reported by the same author at the same 90 days of harvesting age and with the fertilizer of the current study. This difference might become from the genetic variations of grasses and environmental factors (rainfall, temperature, altitude, moisture conditions, light intensity, soil type, and fertility) (Campos et al., 2013) where both studies were conducted.

4.2.7. Crude protein yield

The interaction effects between fertilizer, soil type and type of cultivars on crude protein yield ton per hectare (CPY, t/ha) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.3). The highest CPY was recorded from hybrid Mulato-II (2.43) at black soil with fertilizer followed by B.mutica (2.2) at black soil with fertilizer. The intermediate CPY was from B.mutica (1.77) at red soil with fertilizer, hybrid Mulato- II (1.75) at red soil with fertilizer followed by hybrid Mulato-I (1.46 and 1.41) at red and black soil respectively. While the lowest CPY was from La Liberated (0.28 and 0.22) without fertilizer at red and black soil respectively. The overall mean of CPY was (0.96). Generally, the CPY increased in the current result as dry matter yield increases in all experimental Brachiaria grass cultivars. This might be CPY was obtained by multiplication of dry matter yield by crude protein content.

Fertilizer, soil type, and type of cultivars had a significant interaction effect on CPY of Brachiaria grass in the current study. Generally, the CPY was significantly higher at fertilizer of which high dry matter yield and high CP in all experimental cultivars at both soil types. The current result is in line with the result of desho grass as reported by (Biniyame Mihret et al., 2018) the CPY of desho grass affected significantly by fertilizer type and level of fertilizer. The highest CPY content was observed at NPS fertilizer (2.76) followed by manure application (1.39) while the lowest observed at without fertilizer (0.7) ton per hectare. The current result from all experimental cultivars and overall interaction mean was lower than desho grass as reported by the same author at NPS fertilizer and comparable with the remaining treatment groups. This difference might become from genetics, environmental conditions, soil type and fertility, altitude, interaction effect and harvesting age where both studies were conducted.

79

Similarly, type of cultivars, fertilizer, and cultivars had significant interaction effect on CPY Brachiaria grass in the current study is in line with Wubetie Adnew et al. (2018b) reported that the highest CPY of Brachiaria brizantha grass was recorded from Eth.13809 (0.66 t/ha) followed by Eth.1377 (0.52 t/ha) and Eth.13726 (0.38 t/ha). The same author reported that the highest CPY was recorded from lowland (0.71 t/ha) followed by midland (0.52 t/ha) and highland (0.35 t/ha). Overall, the current interaction result was significantly higher CPY than as reported by the same author except for Brachiaria La Liberated cultivar without fertilizer. This difference might become from genetic potentials of these cultivars on dry matter yield production and CPY content of grasses and environmental conditions as reported by (Ashagre Abate, 2008) where both studies were conducted.

4.2.8. In vitro dry matter digestibility (IVDMD)

The interaction effects between fertilizer, soil type and type of cultivars on in vitro dry matter digestibility (IVDMD %) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.3). The highest IVDMD was from hybrid Mulato-I (68.77 and 68.45) with fertilizer at black and red soil respectively followed by the same cultivar (67.72) without fertilizer at both soil types and Marandu (66.98) at red soil with fertilizer. The intermediate IVDMD was from Mulato-II (65.82) with fertilizer at red soil, hybrid Mulato-II (65.82) at black soil without fertilizer and Marandu (64.41) at red soil without fertilizer. While the lowest IVDMD was recorded from B.mutica (47.27) at red soil without fertilizer. The overall mean of IVDMD was (60.68). In vitro dry matter digestibility is one of many factors influencing animal productivity (Mathison et al., 1995). This influenced by the availability of the degradable materials of the feed nutrients.

The application of fertilizer, soil type, and type of cultivars had a significant (p<0.001) interaction effect on the IVDMD of Brachiaria grass. Significantly higher and lower (p<0.001) IVDMD was obtained from hybrid Mulato-I (68.77) with fertilizer and B.mutica (47.27) without fertilizer respectively. The current interaction result was in line with Brachiaria grass species as reported by (Mutimura et al., 2017) IVDMD affected by the type of species and cultivars. The highest IVDMD was from hybrid Mulato-I (49.5), Marandu (46.5) and hybrid Mulato-II (45.7) at 90 days of harvesting. The current result was higher than as reported by the same author at the same harvesting age within the same species of Brachiaria grass cultivars. This difference might be coming from environmental

80

conditions, soil type and fertility, management system, interaction effect and type of fertilizer were the current experiments were conducted.

Similarly, the current finding is also in agreement with Mustaring et al. (2014) who reported that IVDMD has significantly differed among the type of Brachiaria species. The highest IVDMD was from B.brizantha (57), hybrid Mulato-I (56) and B.mutica (50) at 60 days of harvesting. This was lower than the current result of overall mean at the same 90 days of harvesting age. Susan et al. (2015) reported that IVDMD was affected by the species of Brachiaria grass. The highest IVDMD was from hybrid Mulato-II (51.4), B. decumbens (50.1), La liberated (43.6) and Marandu (28.8) at 90 days of harvesting age. The current result was significantly higher at the same 90 days of harvesting and the same cultivars as reported same author. This difference might become from environmental conditions, interaction, soil type and fertility, management system, and type of fertilizer where both studies were conducted.

4.3.9. In vitro organic matter digestibility (IVOMD)

The interaction effects between fertilizer, soil type and type of cultivars on in vitro organic matter digestibility (IVOMD %) of Brachiaria grasses in the current study was highly significant (p<0.001) (Table 4.3). The highest IVOMD was from hybrid Mulato-I (60.5 and 60.56) with fertilizer at black and red soil respectively followed by the same cultivar (58.77) without fertilizer at both soil types, Marandu (58.34) at red soil with fertilizer and hybrid Mulato-II (57.24) at black soil with fertilizer. The intermediate IVOMD was recorded from La Liberated (52.2 and 52.54) at black soil with and without fertilizer respectively, La Liberated (53.37) at red soil with and without fertilizer, Marandu (55.36) at red soil without fertilizer and hybrid Mulato-II (55.9) at red soil with fertilizer. The lowest IVOMD was from B.mutica (46.5 and 40.66) without fertilizer at black and red soil respectively, hybrid Mulato-II (45.5) at black soil without fertilizer and B.mutica (46.15) at red soil without fertilizer. The overall interaction mean of IVOMD was (52.72).

The application of fertilizer, soil type, and type of cultivars had significant (p<0.001) interaction effect on the IVOMD of Brachiaria grass was observed in the current study. Significantly higher and lower (p<0.001) IVOMD was obtained from hybrid Mulato-I (60.56) and Mutica (46.15) respectively. The overall mean of IVDMD of the current result

81

was (52.72). The current result is in line with Mutimura et al. (2017) reported that IVOMD affected by species and cultivars. The highest IVOMD was from hybrid Mulato-II (50.9), hybrid Mulato (49.4) and Marandu (43.7) at 90 days of harvesting age. The current result was higher than as reported same author at the same 90 days of harvesting age and the same species of Brachiaria grass cultivars. This difference might be coming from environmental conditions, soil type and fertility, management system, and type of fertilizer as well as interaction effects where both studies were conducted.

Similarly, the current result is also in agreement with Mustaring et al. (2014)) who reported that IVOMD was affected by the type of Brachiaria species. The IVOMD was from B. mutica (47), hybrid Mulato-I (58), Brachiaria brizantha (57) at 60 days of harvesting age. This was lower and higher than from the current interaction result of with same cultivars of B.mutica and hybrid Mulato-I respectively. The difference might be coming from environmental conditions, soil type, and fertility, management system, harvesting age and type of fertilizer. Susan et al. (2015) reported that IVOMD differed among species of Brachiaria grass. The highest IVOMD was obtained from hybrid Mulato-II (53.3), B.decumbens (52.3), La Liberated (46.4) and Marandu (46) at 90 days of harvesting age. Bimrew Asmare et al. (2017) reported that IVOMD desho grass was not significantly affected by harvesting day but linearly as harvesting increases the IVOMD was decreased from 90 (45.62) to 120 (43.37) to 150 ( 42,85) harvesting days. This result was lower than from the current interaction result at the same 90 days of harvesting age, due to genetic variation of grass species, environmental conditions, interaction effects and proportion of cell wall content (Erdem et al., 2015).

4.2.10. Metabolizable energy

The interaction effects between fertilizer, soil type and type of cultivars on Metabolizable energy (ME MJ/kg) of Brachiaria grasses in the current study were highly significant (p<0.001) (Table 4.3). The highest ME was recorded from hybrid Mulato-I (8.78 and 8.82) with fertilizer at black and red soil respectively followed by same cultivar (8.56) without fertilizer at both soil types, Marandu (8.36) at red soil with fertilizer and hybrid Mulato-II (8.2 and 8.18) with fertilizer at black and red soil respectively. The intermediate ME was recorded from La Liberated (7.62 and 7.57) at red soil without and with fertilizer respectively, La Liberated (7.52 and 7.6) at black soil with and without fertilizer

82

respectively, Marandu (7.79 and 7.23) at red soil without fertilizer and black soil with fertilizer respectively and hybrid Mulato-II (7.61) at red soil without fertilizer. The lowest ME was from Mutica (6.28) at red soil without fertilizer. The overall interaction means of ME was (7.73). Lonsdale (1989) reported that, feeds having greater than 12.0 MJ/kg of dry matter of metabolizable energy are considered as high-energy content and feeds with less than 9.0 MJ/kg dry matter of metabolizable energy are categorized as low content of energy in the feeds. Hereafter, according to this classification, all Brachiaria cultivars grown in both soil types, as well as both fertilizers in the current study, were as low energy feed content classification, which is below the minimum maintenance energy requirement of livestock production (McDonald et al., 2010).

Types of cultivars, fertilizer and soil type, in the current study had significant effect on the ME of Brachiaria. This result is in line with Mutimura et al. (2017) who reported that ME of hybrid Mulato-I (8), La liberated (8.1) Brachiaria decumbens (8.4), hybrid Mulato-II (7.7) and Marandu (6.5) at 90 days of harvesting age. The current result was higher at the same harvesting age and the same cultivars except Brachiaria La Liberated. This difference could due to variations management system, altitudes, soil type and fertility, environmental conditions where both studies were conducted. On the other hand, Bimrew Asmare et al. (2017) reported that ME desho grass was not significantly affected by harvesting day but, linearly as harvesting increases the ME was decreased from 90 (6.48 to 120 (6.19) to 150 ( 5.87) harvesting days. The overall mean of ME from the current study was (7.73) was higher than desho grass as reported by the same author at the same 90 of harvesting age. This difference might come from genetic variations of grass species, altitudes, soil type, fertility, environmental conditions.

83

Table 4. 3 Effect of fertilizer, soil type and cultivars on crude protein, crude protein yield, digestibility, and metabolizable energy Brachiaria grasses

Variables Parameters ( Dry matter basis) ST FER Cultivars CP (%) CPY(t/ha) IVDMD (%) IVOMD (%) ME (MJ/kg) With Mutica 10.78f 2.2b 55.25g 48.32ghi 7.27ef Mulato-II 13.04bc 2.43a 65.82abc 57.24ab 8.21bc

Mulato-I 12.7cd 1.41d 68.77a 60.5a 8.78a Marandu 13.78a 0.98e 57.1fg 49.35fgh 7.17ef

La liberated 11.6e 0.66h 60.42def 52.2def 7.52de Black Without Mutica 7.2k 0.5i 52.24h 46.15hi 7.12ef Mulato-II 8.65j 0.96e 52.36h 45.5i 6.92f

Mulato-I 9.2h 0.49i 67.72ab 58.78ab 8.56ab Marandu 9.97g 0.38k 57.57efg 49.99efg 7.3def

La liberated 8.03j 0.22m 60.85de 52.54cdef 7.57de Red With Mutica 10.3g 1.77c 52.24h 46.15hi 7.12ef Mulato-II 13.21bc 1.75c 64.35bc 55.91bc 8.17bc Mulato-I 12.6d 1.46d 68.45a 60.56a 8.82a Marandu 14.02a 0.67h 66.98ab 58.34ab 8.36ab La liberated 11.31e 0.66h 62.6cd 53.37cde 7.62de

84

Table 4.4 (continued) Out Cultivars CP (%) CPY(t/ha) IVDMD (%) IVOMD (%) ME (MJ/kg) Mutica 7.1k 0.43jk 47.27i 40.7j 6.28g Mulato-II 9.02hi 0.74g 58.89ef 51.37efg 7.61de Mulato-I 8.8i 0.5i 67.72ab 58.78ab 8.56ab Marunda 10.7f 0.84k 64.41bc 55.36bcd 7.79cd La liberated 8.19j 0.28l 62.59cd 53.36cde 7.62de Overall Mean 10.55 0.96 60.68 52.72 7.72 SEM 0.34 0.1 1.01 0.89 0.11 CV 1.62 2.58 2.55 3.00 2.72 R2 0.997 0.999 0.979 0.959 0.952 Sig *** *** *** *** ***

ST= Soil type, FER= Fertilizer, CP= Crude protein, CPY= Crude protein yield, IVDMD= In vitro dry matter digestibility, IVOMD= In vitro organic matter digestibility, ME = Metabolizable energy, R2= Coefficient of determination, CV= Coefficient of variation, SEM= Standard error of mean, Sig= Significance level

85

4.3. Farmers’ perception and cultivar selection of Brachiaria grass cultivars

The perception of farmers on Brachiaria grass cultivars selection at both soil types were presented in (Table 4.4, 4.5 and 4.6). Pair-wise ranking of the farmer’s selection criteria was made to rank the selection criteria and to identify the most important trait for the community for future forage development. Participant farmers were using pair-wise ranking procedure at both soil type and identified about five cultivars selection criterions (Table 4.4 and 4.5). Those selection criterions were plot cover (visual observation of the number of plants grown in that area-reflects survival rate), number of tillers (visual observation on the number branches on the specific area), Plant height (visual observation on the plant growth situation), leafiness (visual observation on the number of leaves per plant) and smoothnes (observation made by touching leaves by using fingers). Participant farmers in selection of Brachiaria grass cultivars were voluntary to compare the criteria and rank them in order of importance.

According to pair wise ranking matrix: plot cover, number of tillers, plant height, leafiness, and leaf smoothness criterions were ranked in first, second, third, fourth and fifth respectively (Table 4.4) at black soil while number of tillers, leafiness, plant height, plot cover and smoothness criterions were ranked in first, second, third, fourth and fifth respectively (Table 4.5) at red soil. The value for each selection criterion was defined and provided different points (scores) according to the rank of each selection criterion. Five points, four points, three points, two points and one point for first, second, third, fourth and fifth selection criterion respectively, on their order of importance (De Boef, W.S. and M.H. Thijssen, 2007). All participant farmers’ were agreed on those selection criterions and they might provide a different rank for each selection criterion to select specific Brachiaria grass cultivars at both soil types. As a result, the total weighted score value for each selection criterion was computed as the point provided based on its rank multiplied by the total participant farmers’ vote given for each selection criterion. While the total points for each cultivar were obtained as the sum of the point of all selection criterion (Table 4.7, Appendix 7. 1.5, 7.1.6).

According to the preference ranking, hybrid Mulato-II had the highest weighted value or score (341 and 322) followed by hybrid Mulato-I (326 and 273) at black and red soil respectively. The intermediate weighted score was recorded from Mutica (302 and 214) at

86

black and red soil respectively. Consequently, based on result of preference ranking hybrid Mulato-II , hybrid Mulato-I and B. mutica cultivars were selected in the first, second and third order of ranking respectively at both soil types. The Marandu and La liberated cultivars had lower score according to their preference ranking of which, they selected these cultivars in the fourth and fifth rank respectively at both soil type. This selection of Brachiaria grass cultivars by farmers’ were depending on qualitative parameters such as plot cover, number of tillers, plant height, leafiness and leaf smoothnes for black soil while number of tillers, leafiness, plant height, plot cover and smoothnes for red soil type.

Table 4. 4 Pair-wise ranking matrix of selection criteria at black soil

Weig Plot Plant Tiller Smoothne To Ra Criteria Leafiness hted cover height number s tal nk value Plot Plot Plot cover Plot cover Plot cover * cover cover 4 1 5 Plant Plant Tiller Plant 3 * 2 3 height height number height Tiller 2 Leafiness * Leafiness 1 4 number Tiller Tiller 4 * 3 2 number number Smoothnes * 0 5 1

Table 4. 5 Pair-wise ranking matrix of selection criteria at red soil

Weig Plot Plant Leafines Tiller Tot Ra Criteria Smoothnes hted cover height s number al nk value Plant Leafines Tiller Plot cover Plot cover * height s number 1 4 2 Plant Leafines Tiller Plant 3 * 2 3 height s number height Tiller 4 Leafiness * Leafiness 3 2 number Tiller Tiller 5 * 4 1 number number Smoothnes * 0 5 1

87

Table 4. 6 Farmers’ preference ranking of Brachiaria grass cultivars

selection Weighte At black soil or Andassa district criteria d value Brachiaria grass cultivars and weighted score values B. mutica Mulato-II Mulato-I Marandu La Liberated Plot cover 5 125(25) 120(24) 120(24) 65(13) 70(14) Tiller number 4 72(18) 80(20) 88(22) 64(16) 56(14) Plant height 3 60(20) 60(20) 69(23) 51(17) 15(5) Leafiness 2 30(15) 56(28) 24(12) 60(30) 30(15) Leaf Smoothnes 1 15(1) 25(1) 25(1) 20(1) 10(1) Total score - 302 341 326 260 181 Rank - 3 1 2 4 5 At red soil or North Mecha district Tiller number 5 70(14) 110(22) 85(17) 75(15) 45(9) Leafiness 4 56(14) 104(26) 60(15) 48(12) 52(13) Plant height 3 48(16) 48(16) 60(20) 36(12) 18(6) plot cover 2 24(12) 44(22) 52(26) 16(8) 24(12) Leaf Smoothnes 1 16(1) 16(1) 16(1) 22(1) 10(1) Total score - 214 322 273 197 149 Rank - 3 1 2 4 5

4.4. Partial Budget Analysis

The overall partial budget analysis of the current study was presented at both soil type in (Table 4.7 and 4.8). The highest net benefit in the black soil was obtained from B. mutica agrass cultivar (53447.9birr) followed by hybrid Mulato-II (48648.47birr). The intermediate net benefit was recorded from hybrid Mulato-II grass cultivar (30187.89 birr) while the lowest was recorded from Brachiaria brizantha La Liberated (7499.25birr) without fertilizer application (Table 4.7). Similarly the highest net benefit in the black soil was obtained from B. mutica grass cultivar (44877.13birr) followed by hybrid Mulato-II (34241 birr). The intermediate net benefit was recorded from hybrid Mulato-I grass cultivar (29396 birr) while the lowest was recorded from Brachiaria brizantha La Liberated (9462.7) without fertilizer application (Table 4.9). According to dominance analysis, all cultivars except Brachiaria mutica and hybrid Mulato-II at black and red soil respectively without

88

fertilizer application were dominated by other treatments with fertilizer application, hence, eliminated from further economic analysis. Based on the marginal analysis, Brachiaria mutica (1641.85 and 1066.85%), hybrid Mulato-II (1413.41 and 565.42%) and hybrid Mulato-I (438.998 and 334.82% were superior marginal rate of return (MRR) to other treatments at black and red soil respectively (Table 4.7 and 4.8).

The net benefits among cultivars, fertilizer, and soil type were quite different. This is due to the production potential of dry matter yield differences among experimental Brachiaria grass cultivars at both fertilizer and soil type. As result of current finding, the marginal rate of return obtained from Brachiaria Mutica (1641.85 and 1066.85 % MRR)), hybrid Mulato- II (1413.41 and 565.42% MRR) and hybrid Mulato-II (438.998 and 334.82 % MRR), implies that for one Birr investment on Brachiaria grass production, the producer can get Birr (16.4285 and 10.6685), (14.1341 and 5.652) and (4.38998 and 3.3482) at black and red soil respectively. The current MRR on these cultivars were recorded above the minimum acceptable rate of return (CIMMYT, 1988). Generally, at both soil types, fertilizer had more responses to high net benefits and MRR of B.Mutica, hybrid Mulato-II and hybrid Mulato- I in the current study without considering other benefits and costs that are fixed throughout all treatment groups, so this partial budget analysis was only considering costs that vary at experimental cultivars throughout the experiment period as procedures (CIMMYT, 1988).

89

Table 4. 7 Partial budget analysis Brachiaria grass cultivars grown at black soil with and without fertilizer

Description Brachiaria grass cultivars (Treatments ) La Marandu Mulato-I Mulato- Mutica La Marandu Mulato-I Mulato- Mutica Liberated (WO) (WO) II (WO) (WO) Liberated (WF) (WF) II (WF) (WF) (WO) (WF) DMY (Kg) 2750 3850 5330 6500 6950 5700 7130 11070 18610 20370 ADMY (Kg) (A) 2475 3465 4797 5850 6255 5130 6417 9963 16749 18333 Cost of NPS 0 0 0 0 0 1476 1476 1476 1476 1476 Cost of Urea 0 0 0 0 0 625 625 625 625 625 (TVC) (B) 0 0 0 0 0 2101 2101 2101 2101 2101 Price (birr/ kg) (C) 3.03 3.03 3.03 3.03 3.03 3.03 3.03 3.03 3.03 3.03 Gross benefit (A*C)(D) 7499.25 10498.95 14534.91 17725.5 18952.65 15543.9 19443.51 30187.89 50749.47 55548.99 Net benefit (D-B)(birr) 7499.25 10498.95 14534.91 17725.5 18952.65 13442.9 17342.51 28086.89 48648.47 53447.99 Change in TVC (∆TV) 0 0 0 0 0 2101 2101 2101 2101 2101 Change in net benefit 7499.25 10498.95 14534.91 17725.5 0 -5509.75 -1610.14 9223.35 29695.82 34495.34 (∆NR) MRR (%) = (∆NR/∆TV ND ND ND ND ND D D 438.998 1413.41 1641.85 Cost)*100

WO= without fertilizer, WF=with fertilizer, ND= Not determine, D=dominated, DMY= Dry matter yield, ADMY= Adjusted dry matter yield, NPS= Nitrogen, phosphorous and sulfur, MRR= Marginal rate of return, TVC=Total variable cost 90

Table 4. 8 Partial budget analysis Brachiaria grass cultivars grown at red soil with and without fertilizer

Description Brachiaria grass cultivars ( Treatments ) La Marand Mutica Mulato-I Mulato- La Marandu Mulato-I Mulato-II Mutica Liberated u (WO) (WO) (WO) II (WO) Liberated (WF) (WF) (WF) (WF) (WO) (WF) DMY (Kg) 3470 4440 5270 5660 8200 5870 6020 11550 13290 17190 ADMY (Kg) (A) 3123 3996 4743 5085 7380 5283 5418 10395 11961 15471 Cost of NPS 0 0 0 0 0 1476 1476 1476 1476 1476 Cost of Urea 0 0 0 0 0 625 625 625 625 625 (TVC) (B) 0 0 0 0 0 2101 2101 2101 2101 2101 Price (birr/kg) (C) 3.03 3.03 3.03 3.03 3.03 3.03 3.03 3.03 3.03 3.03 Gross benefit (A*C)(D) 9462.7 12107.9 14371.3 15407.55 22361.4 16007.49 16416.54 314496.85 36241.83 46877.13 Net benefit (D-B) 9462.7 12107.9 14371.3 15407.55 22361.4 13096.49 14316.54 29396 34241 44776 Change in TVC (∆TV) 0 0 0 0 0 2101 2101 2101 2101 2101 Change in net benefit 9462.69 12107.9 14371.3 15407.55 22361.4 -9264.91 -8044.86 7034.6 11879.6 22414.6 (∆NR) MRR (%) = (∆NR/∆TV ND ND ND ND ND D D 334.82 565.42 1066.85 Cost)*100

WO= without fertilizer, WF=with fertilizer, ND= Not determine, D=dominated, DMY= Dry matter yield, ADMY= Adjusted dry matter yield, NPS= Nitrogen, phosphorous and sulfur, MRR= Marginal rate of return, TVC=Total variable cost 91

4.5. Correlation among morphology, dry matter yield and quality parameters of Brachiaria grass cultivars

Simple linear bivariate correlation analysis among morphological, forage yield, nutritive value parameters of Brachiaria grass cultivars were presented in (Table 4.9). The plant height was positively correlated to LLPP, NLPP (p > 0.05), DMY, CPY, NDF, ADF and ADL (p > 0.01), but negatively correlated with Ash (p > 0.01), LSR, IVDMD and IVOMD (p > 0.05). The number of tillers and the number of leaves per plant were positively correlated (p > 0.01) to DMY, CPY and CP (p > 0.05) and with each other. The LSR was positively (p > 0.01) correlated to Ash, CP, IVDMD, IVOMD, and ME but negatively correlated with DM content (p>0.05), NDF and ADF ((p >0.01). The DMY was positively correlated with NLPP, PH, NTPP, ADL (p>0.05) and CPY (p>0.01).

The plant height was a moderate correlation with most growth parameters, DMY, CPY, fiber contents but negative correlation with LSR and Ash in the current study is in line with Brachiaria Mutica grass as reported by (Mimilia Zemene, 2018). The number of tillers and number of leaves per plant was a positively strong correlation with DMY, CPY and positive moderate correlation with CP and with each other in the current study are in line with Brachiaria Mutica grass as reported the same author. The positive correlations of most morphological parameters with each other and negative correlation with LSR in the current study are in line with desho grass as reported by (Bimrew Asmare et al., 2017, Genet Tilahun et al., 2017).

The LSR was a positively strong correlation with CP and moderate with Ash, IVDMD, IVOMD, and ME, but moderate negative correlation with DM content, NDF, and ADF in the current study. This is in line with Brachiaria Mutica as indicated by Mimilia Zemene, 2018) on CP and Ash parameters and with desho grass by Bimrew Asmare et al., 2017) reported on IVOMD and ME. This showed that LSR is an important factor associated with the digestibility of forage feeds by ruminant animals (Yasin et al., 2003). Tessema Zewdu et al. (2002) also previously observed the direct relationship between leaf-to-stem ratio and CP content, IVDMD, IVDMD and the inverse association of leaf-to-stem ratio and fiber content for Napier grass. The DMY was a strong positive correlation with CPY and a moderate positive correlation with most growth parameters in the current study is in line with Brachiaria Mutica and desho grass as indicated by (Mimilia Zemene, 2018 and by

92

Bimrew Asmare et al., 2017) respectively. This showed that DMY had a direct relationship with growth parameters and CPY in the current study. Similarly, the DMY positive correlation with PH, NLPP, LLPP and negative correlation with ASH in the current study is in line with desho grass as indicated by (Genet Tilahun et al., 2017).

The DM content was significantly (p > 0.05) negatively correlated to CP and LSR. The ASH content was significantly (p > 0.01) negatively correlated to PH, LSR, NDF ADF, and ADL but positively correlated to IVDMD and IVOMD. The fiber contents negatively correlated to IVDMD, IVOMD and ME, but positively correlated with each other. The CP content was significantly (p > 0.05) positively correlated to CPY, IVDMD, IVOMD, ME, NTPP, NLPP and LSR (p > 0.01) but negatively correlated with DM and fiber content parameters (p > 0.01). The CPY was significantly (p > 0.01) positively correlated to PH, NTPP, NLPP, DMY and CP (p > 0.05). The IVDMD, IVOMD and ME significantly positively correlated to LSR, ASH, CP, and with each other and negatively correlated to PH and fiber parameters. This is the fact that as plant height increased the nutritive value parameters related to digestibility and CP were decreased and vice versa.

The Ash content was a moderate negative correlation with growth parameters in the current study is disagreement with previous reports of with Bimrew Asmare et al. (2017) and Biniyame Mihiret et al. (2018) of desho grass which was no correlation with this parameters. This might come from genetic and environmental variations where both studies were conducted. The fiber parameters (NDF, ADF, and ADL) contents were strong positive correlation with each other and strong negative correlation with IVDMD, IVOMD and CP content is in the current study in line with desho grass as reported by Bimrew Asmare et al., 2017; Biniyame Mihiret et al.,2018) on the CP and fiber content parameters. This is due to the fact that as the fiber content of forage decreases the digestibility and CP was increased and vice versa. The IVDMD, IVOMD, ME and CP contents were significantly positively correlated with each other and negatively correlated with fiber contents.

93

Table 4. 9 Correlation among on morphological parameters, dry matter yield and quality of Brachiaria grass cultivars

PH LL TN NL LSR DMY DM ASH NDF ADF ADL CP CPY IVDMD IVOMD ME PH 1 0.5* 0.3 0.5* -0.54* 0.75** 0.29 -0.7** 0.63** 0.73** 0.74** -0.04 0.6** -0.54* -0.49* -0.39 LL 1 0.4 0.37 0.07 0.43 -0.21 -0.03 0.05 0.2 0.28 0.32 0.44 -0.08 0-.075 -0.08 TN 1 0.87** 0.36 0.73** -0.17 0.02 -0.28 -0.16 0.2 0.54* 0.79** 0.34 0.37 0.42 NL 1 0.26 0.85** -0.1 -0.05 -0.19 -0.11 0.18 0.53* 0.9** 0.25 0.28 0.33 LSR 1 -0.02 -0.52* 0.62** -0.74** -0.7** -0.42 0.77** 0.17 0.71** 0.7** 0.63** DMY 1 0.13 -0.35 0.1 0.22 0.47* 0.42 0.96** -0.03 0.024 0.12 DM 1 -0.37 0.22 0.15 0.14 -0.46* -0.03 -0.09 -0.07 0.02 ASH 1 -0.62** -0.8** -0.73** 0.41 -0.19 0.63** 0.56** 0.39 NDF 1 0.85** 0.61** -0.47* -0.07 -0.81** -0.8** -0.73** ADF 1 0.81** -0.41 0.04 -0.84** -0.81** -0.71** ADL . 1 -0.06 0.04 -0.54* -0.47* -0.34 CP 1 0.56* 0.51* 0.53* 0.49* CPY 1 0.09 0.13 0.2 IVDMD 1 0.99** 0.96** IVOMD 1 0.98** ME 1

Level of significance: ** = P<0.01; * = P<0.05, PH = plant height, TN = number of tillers per plant, LL = leaf length, LSR = Leaf to Stem Ratio, NL = number of leaves per plant, DMY = dry matter yield, CPY = crude protein yield , DM = dry matter, Ash=ash content, NDF = neutral detergent fiber, ADF = Acid detergent fiber, ADL = acid detergent lignin , CP = crude protein, IVDMD = in vitro dry matter digestibility, I OMD = in vitro organic matter digestibility, ME = metabolizable energy 94

Chapter 5. CONCLUSION AND RECOMMENDATIONS

5.1. Conclusion

The interaction effects of fertilizer, soil type and different Brachiaria grass cultivars for morphological characteristics, dry matter yield, and nutritive value observed in this study. The application of fertilizer had a significant effect on the morphological characteristics, dry matter yield and chemical composition of Brachiaria grass cultivars at both soil types as compared to without fertilizer. The effect of soil type on morphological characteristics, dry matter yield and chemical composition of Brachiaria grass cultivars were observed in which it is better at black soil than red soil at both with and without fertilizer.

In most morphological characteristics and dry matter yield, Brachiaria mutica cultivar had better yield at both soil types followed by hybrid Mulato-II and hybrid Mulato-I. The cultivars of hybrid Mulato-I, hybrid Mulato-II and Marandu cultivars were a better nutritional value like high CP, IVDMD, IVOMD and lower NDF, ADF and ADL contents at both soil types with fertilizer as compared to without fertilizer. Brachiaria mutica cultivar was the lowest nutritive value and the highest fiber content of all experimental cultivars. In the farmers’ perception by cultivar selection, hybrid Mulato-II was selected followed by Mulato-I and Brachiaria mutica an in ranking order, respectively. Moreover in the partial budget analysis, the higher net benefit and MRR was obtained from Brachiaria mutica followed by hybrid Mulato-II and hybrid Mulato-I at both soil types. The correlation between PH, LLPP, NLPP, DMY, CPY, NDF, ADF, and ADL was positive but negatively correlated with Ash, CP LSR, IVDMD, and IVOMD.

Generally, CP content obtained from all Brachiaria cultivars in the current study can fulfill the minimum requirements of a ruminant animal which ranges from (10 to 14%) with fertilizer. While metabolizable energy obtained from all experimental Brachiaria grass cultivars in the current study was a below minimum maintenance requirement of ruminant animals. Based on the all aspects of evaluations, Brachiaria hybrid Mulato-II, Brachiaria Mutica and Brachiaria hybrid Mulato-I, cultivars were selected as adaptive and better production performance to fulfill the forage quantity and quality shortage to enhance the livestock production and productivity.

95

5.2. Recommendations

Based on the current result and above conclusion the following suggestions were forwarded for future researches, wider adaptations and promotion of this climate-smart Brachiaria grass cultivars in West Gojam, Ethiopia. Therefore, having on this general information about these Brachiaria grass cultivars beneficiary making decisions based on the relative significance of forage dry matter yield, nutritional nutritive value, farmers' perception and net benefit in farmers' practical farming systems, the following recommendations are forwarded.

 Among the tested cultivars, Brachiaria hybrid Mulato-II, Brachiaria Mutica and Brachiaria hybrid Mulato-I cultivars are recommended alternatively for wider adaptation and on-farm evaluation in study areas, similar soil types and agro-ecological zones

 Application of fertilizer on Brachiaria grass cultivars economically advantageous than without fertilizer in terms of production and forage nutritive value, hence the application of fertilizer is recommended for better dry matter yield and nutritional nutritive value at both soil types.

 To fully utilize the potentials of Brachiaria grass species and cultivars, further studies on agronomic and nutritional evaluations involving live-animal experiments are recommended

96

Chapter 6. REFERENCES

Abdi Hassan, Tessema Zewdu, Mengistu Urgeand Sisal Fikru. 2015. Effect of Nitrogen Fertilizer Application on Nutritive Value of Cenchrus ciliaris and Panicum Maximum Grown under Irrigation at Gode, Somali Region.J Nutr Food Sci 2015, S11:005 DOI: 10.4172/2155-9600.1000S1100.

Abdullah, A.S. 2015. Natural Toxicants Affecting Animal Health and Production; Available online

Abebe Mekoya. 2008. Multipurpose fodder trees in Ethiopia; Farmers’ perception, constraints to adoption and effects of long-term supplementation on sheep performance. PhD thesis, Wageningen University, Wageningen, the Netherlands.

Abule Ebro, 2015. Changes in grazing land management and implications on livestock production in west Shoa zone, international livestock research institute, Ethiopia.

Akman and Topal, .2014. Effects on Wheat Root Growth and Development of Rhizosphere Environment and Stress Factors. Journal of Seleuk University. Natural and Applied science http://josunas.selcuk.edu.tr/login/index.php/josunas/about.ISSN: 2147-3781.

Alemayehu Mengistu and Getnet Assefa, 2012. Animal feed resource research in Ethiopia: achievement

Alemayehu Mengistu, Gezahagn Kebede, Fekede Feyissa and Getnet Assefa. 2017. Review on Major Feed Resources in Ethiopia: Conditions, Challenges and Opportunities. Academic Research Journal of Agricultural Science and Research.Vol. 5(3), pp. 176- 185, May 2017.

Alemayehu Mengistu. 2003. Country pasture/forage resources profiles: Ethiopia. FAO web site: http://www.fao.org/ag/agp/agpc/doc/counprof/ethiopia/ ethiopia.htm.

Alemayehu Mengistu. 2004. Pasture and forage resource profiles of Ethiopia. EDM printing press, Addis Ababa, Ethiopia.

Alemayehu Mengistu. 2006a. Country Pasture/Forage Resource Profiles, Ethiopia. Available at: http://www.fao.org/ag/agp/agpc/doc/counprof/ethiopia/ethiopia.htm.

Alemayehu Mengistu. 2012. Forage Seed Research and Development in Ethiopia. Ethiopia Institute of Agricultural Research, Addis Ababa, Ethiopia. Pp 3-5.

Alemu Yami. 2009. Nutrition and feeding of sheep and goat. In: Ethiopian sheep and goat production improvement program, Pp: 106-159.

97

Almeida, C.M., Lana, Â.M.Q., Rodrigues, J.A.S., Alvarenga, R.C. and Borges, I. 2012. Influência do Tipo de Semeadura a Produtividade do Consórcio Sorgo—Urochloa brizantha CV Marandu no Sistema de Integração Lavoura-Pecuária.Revista Brasileira de Milho e Sorgo, 11, 60-68.

ALRC. 2018. Participatory evaluation of productive Napier accessions in north western Amhara Region Ethiopia. Presentend at Annual completed review forum, Amhara Agreicultural Research Institute (Unpublished).

Ammar H, López S and Andrés S. 2010. Influence of maturity stage of forage grasses and leguminous on their chemical composition and in vitro dry matter digestibility. In: Porqueddu C. (ed.), Ríos S. (Ed.).The contributions of grasslands to the conservation of Mediterranean biodiversity. Zaragoza: CIHEAM / CIBIO / FAO / SEEP, 2010. P199- 203.

AOAC. 2004. Official methods of analysis. (16th Ed.) Association of Official Analytical Chemists, Washington DC.

Arango J., Moreta D., Núñez J., Hartmann K., Domínguez M., Ishitani M., Miles J., Subbarao G., Peters M., Rao I. 2014. Developing methods to evaluate phenotypic variability in biological nitrification inhibition (BNI) capacity of Brachiaria grasses. Tropical Grasslands—Forrajes Tropicales, 2: 6-8.

Arelovich HM, Abney CS, Vizcarra JA, Galyean M. 2008. Effects of dietary neutral detergent fiber on intakes of dry matter and net energy by dairy and beef cattle: analysis of published data. Prof Anim Sci. 24:375–383.

Argel, P.J., J.W. Miles, J.D. Guiot, and C.E. Lascano. 2005. Cultivar Mulato (Brachiaria hybrid).

Argel, P.J., Miles, J.W. Guiot, J.D Cuardrado, H. and Lascano, C.E. 2007. Cultivar Mulato II.high nutitive value forage grass, resistant to spittle bug and adapted to a well-drained acid tropical soil. Cali, Colombia: International Centre for Tropical Agriculture (CIAT), 21pp.

Arzani, H., Sadeghimanesh, M.R., Azarnivad, H., Asadian, G.H., Shahriyari, E. 2008. Study of phonological stages effect values of twelve species in Hamadan rangelands. Iran J Range Desert Res. 16:86–95.

Ashagre Abate.2008. Effect of nitrogen fertilizer and harvesting stage on yield and nutitive value of natural pastures in fogera district, north western Ethiopia. Msc thesis shool of graduate studies, Haramaya University.

98

Ball,D.M., M.Collins, G.D. Lacefield, N.P. Martin,D.A. Mertens, K.E.Olson,D.H. Putnam,D.J. Undersander, and M.W.Wolf. 2001. Understanding Forage Nutitive value.American Farm Bureau Publication 1-01, Park Ridge, IL.

Behnke, R. 2010. The Contribution of Livestock to the Economies of IGAD Member States: Study Findings, Application of the Methodology in Ethiopia and Recommendations for Further Work, IGAD LPI Working Paper 02-10. Odessa Centre, IGAD Livestock Policy Initiative, Great Wolford, UK.

Berhanu, Gebremedihn, Adane, Hirpa and Kahsay, Berhanu, 2009. Feed marketing in Ethiopia: Results of rapid market appraisal. Improving Productivity and Market Success IPMS) of Ethiopian farmers project Working Paper 15. ILRI (International Livestock Research Institute), Nairobi, Kenya, pp: 64

Berihun Melke. 2005. Effect of Planting Patterns and Harvesting Days on Yield and Nutitive value of Bana Grass [Pennisetum purpureum (L.)×Pennisetum americanum (L.)] (Master thesis) School of Graduates of Haramya University, Harar, Ethiopia.

Beyadglign Hunegnaw, Wondimeneh Mekonnan, Misganaw Walie, Agraw Amanie and SimagnewTamir. 2016. Participatory Selection of Different Lablab Accessions for Under Sowing on Maize In Western Gojam, Ethiopia. Proceeding of the 9th Annual Regional Conference on Completed Livestock Research Activities on 9-20 , 2015.

Bill DeBoer, 2016. Water and Nutrient Uptake by Roots.

Bimrew Asmare, Solomon Demeke, Taye Tolemariyam, Firew Tegegne, Aynalem Haile, Jane Wamatu. 2017. Effects of altitude and harvesting dates on morphological characteristics, yield and nutritive value of desho grass (Pennisetum pedicellatum Trin.) in Ethiopia.

Biniyname Mihret, Bimrew Asmare, Yeshambel Mekuriaw. 2018. Effect of fertilizer type and plant spacing on plant morphological characteristics, yield and chemical composition of desho grass (Pennisetum pedicellatum Trin.) in Northwestern Ethiopia. Agricultural Science and Technology, VOL. 10, No 2, pp 107 - 114, 2018.

Brandi, R. A., Tribucci, A. M. O., Balieiro, J. C. C., Hoffman, R. H., & Bueno, I. C. S. 2014. Citrus pulp in concentrates for horses. Food and Nutrition Sciences, 5, 1272–1279. doi:10.4236/fns.2014.513138

Cameron, A. G. and B. Lemcke. 2008. Para grass: a pasture grass for wet and flooded soils. No: E30.

Campos, F.P., Sarmento, P., Nussio, L.G., Lugão, S.M.B., Lima, C.G. and Daniel, J.L.P. 2013. Fibre monosaccharides and digestibility of Milenio grass under N fertilizer. Animal Feed Science and Technology 183, 17–21. 99

CASCAPE (Capacity building for scaling up of evidence based best practices in Agriculture production in Ethiopia). 2015. CASCAPE Manual for best fit practice Rhodes grass. BDU-CASCAPE working paper 10.

Chanthakhoun, V, Wanapat, M. and Berg, J. 2012. Level of crude protein in concentrate supplements influenced rumen characteristics, microbial protein synthesis and digestibility in swamp buffaloes (Bubalus bubalis). Livestock Science, 144, 197–204.

Chaves MM, Pereira JS, Maroco J, Rodrigues ML, Ricardo CPP, Osório ML, Carvalho I, Faria T, Pinheiro C .2002. How plants cope with water stress in the field. Photosynthesis and growth. Annals of Botany 89, 907–916

CIAT. Annual Report. 2011. Project IP-5. Tropical Grasses and Legumes: Optimizing genetic diversity for multipurpose use. p. 88-90.

CIAT. Annual Report. 2005. Methodology development for screening Brachiaria hybrids for high digestibility and protein. P. Avila, G. Ramirez, C.E. Lascano, and J. Miles Project IP-5: Tropical gasses and Legumes: Optimizing genetic diversity for multipurpose use. 266p.

CIAT. Annual Report. 2006. Screening of sexual and apomictic Brachiaria hybrids for digestibility and protein. Lascano, C.E., J. Miles, P. Avila, and G. Ramirez. Project IP- 5: Tropical gasses and Legumes: Optimizing genetic diversity for multipart.

CIMMYT (International Maize and Wheat Improvement Center), 1988. Farm Agronomic to farmers’ recommendation. An Economic Training Manaual. Completely revised edition, D.F. Mexico. 51p.

Clara Machogu. 2013. A comparative study of the productivity of Brachiaria hybrid cv. mulato ii and native pasture species in semi-arid rangelands of Kenya. MSc. Thesis a thesis in range management of the University of Nairobi, Kenya.

Cook, B.G.:Pengelly.B.C.:Brown, S.D.: Donnelly, J.L.: Eagles, D.A,:Franco,M.A.: Hanson,J.; Mullen,B.F; Partridge, I.J.; Peters, M,; Schultze-Kraft, R. 2005. Tropical forages CSIRO, DPI$F (Qld), CIAT AND ILRI, Brisbane, Australia.

Crowder, L.V. and Chheda M.R. 1982. Tropical grassland husbandry. Longman, London. p562.

Crush, J.R.; Easton, H.S.; Waller, J.E. 2006. Genetic variation in perennial ryegrass for root profile characteristics. pp. 63-65. In: Advances in pasture plant breeding, 13th Australasian Plant Breeding Conference.

Crush, J.R.; Nichols, S.N. 2010. Progress towards forage plant root systems for sustainable dairying. pp. 50-52. In: Proceedings of the 4th Australasian Dairy Science Symposium.

100

CSA. 2018: Agricultural sample survey 2017/18 [2010 E.C.] Volume II report on livestock and livestock characteristics, Federal Democratic Republic of Ethiopia, Central Statistical Agency; Statistical bulletin number 587.

Danes, M.A.C., Chagas, L.J., Pedroso, A.M. and Santos, F.A.P. 2013. Effect of protein supplementation on milk production and metabolism of dairy cows grazing tropical grass. Journal of Dairy Science, 96, 407–419.

De Boef, W.S. and M.H. Thijssen, 2007. Participatory tools working with crops, varieties and seeds. A guide for professionals applying participatory approaches in agro biodiversity management, crop improvement and seed sector development. Wageningen, Wageningen International, 83pp.

Depeters E.1993. Forage nutitive value and its implications. In: California Alfalfa workshop (Eds) proceedings of the 23rd California alfalfa symposium at Fresno, California, USA, pp. 93-146. Visala, California, USA: Department of agronomy and range sciences extension.

Dewis, J and P. Freitas. 1975. Physical and chemical methods of soil and water analysis.FAO bul. No. 10, Rome. P 275

Diriba Geleti and Adugna Tolera. 2013. Effect of age of regrowth on yield and herbage nutitive value of panicum coloratum under sub humid climatic conditions of Ethiopia. African journal of Agricultural research. Vol. 8(46), Pp. 5841-5844.

El-Nashaar, H.M., Griffith, S.M., Steiner, J.J. and Banowetz, G.M. 2009. Mineral concentration in selected native temperate grasses with potential use as biofuel feedstock. Bio resource Technology, 100, 3526–3531.

Endale Yadessa, Abule Ebro, Lemma Fita and Getnet Asefa, 2016. Feed resources and its utilization practices by smallholder farmers in meta-robi district, west shewa zone, oromiya regional state, Ethiopia.Vol. 4(4), Pp. 124-133.

Erdem .F, N. Cetinkaya, C. Nisbet & E. Altin. 2015. Estimation of organic matter digestibility, metabolizable energy, phenolic compounds and antioxidant activity of stems and seeds of the Juncus acutus plant in ruminant nutrition. South African Journal of Animal Science 2015, 45 (No. 5).

Ethiopian feed resources database. (http://www.vslp.org/ETHFeed/default.asp)

Etsubdink Tekalign .2014. Forage seed systems in Ethiopia: A scoping study. ILRI Project Report. Nairobi, Kenya: International Livestock Research Institute (ILRI).

FAO (Food and Agriculture Organization of the United Nations), 2008. FAO fertilizer and plant nutrition bulletin: Guide to laboratory establishment for plant nutrient analysis. FAO, Rome, Italy. 203p. 101

Firew Tegegne and Getnet Assefa. 2010. Feed source Assessment in Amhara National Region state. Ethiopia sanitary and Phytosanitary Standards and livestock and meat marketing program (SPS-LMM) Texas A and M University System, pp.104.

Fekede Feyissa, Agaje Tesfaye and Angaw Tsigie. 2011. Producing and using alternative feeds to crop residues efficient use of crop residues: Animal Feed Vs Conservation Agriculture Research Project. Ethiopian Institute of Agricultural Research (EIAR). Addis Ababa, Ethiopia

Gebre/Mariam S, Amare S, Baker D, Solomon A, Davies R. 2013. Study of the Ethiopian live cattle and beef value chain. ILRI Discussion paper 23. Nairobi, Kenya; Addis Ababa, Ethiopia: International Livestock Research Institute (ILRI).

Genet Tilahun, Bimrew Asmare and Yeshambel Mekuriaw. 2017. Effects of harvesting age and spacing on plant characteristics, chemical composition and yield of desho grass (Pennisetum pedicellatum trin.) In the highlands of Ethiopia. Vol. 5(2):77–84.

Getnet Assefa and Inger Ledin. 2001. Effect of variety, soil type and fertilizer on the establishment, growth, forage yield, nutitive value and voluntary intake by cattle of oats and vetches cultivated in pure stands and mixtures.Animal Feed Science and Technology 92 (2001) 95 111. © 2001 Elsevier Science.

Ghimire Sita, Donald Njarui, Mupenzi Mutimura, Juan Cardoso, Linda Johnson Elias Gichangi, Suliana Teasdale, Kennedy Odokonyero, John Caradus, Idupulapati Rao and Appolinaire Djikeng. 2015. Climate-smart Brachiaria for improving livestock productionin East Africa: Emerging opportunities.

Gomez, K. A. And Gomez A. A. 1984. Statistical Procedures for Agricultural Research. 2nd Edition. New York, USA: John Wiley and Sons.

Guiot, J.D., and F. Melendez. 2003. Mulato Pasture. Brachiaria hibrid (CIAT 36061). Excellent alternative to meat and milk production in tropical areas. Publication of Tabasco, Semill as Papalotla in ISPROTAB. Villahermosa, Tabasco (Mexico). November, 2003. 23p.

HARC (Holeta Agricultural Research Center). 2004. Feed Resources Development and Utilization: Possible Options and Recommendations under Ethiopia Condition.

Hare, M.D., Tatsapong, P. and Saipraset, K. 2007. Seed production of two Brachiaria hybrid cultivars in north-eastThailand. 3. Harvesting method. Tropical Grasslands, 41, 43–49.

Hazelton, P. and Murphy, B. 2007. Interpreting soil test results: what do all the numbers mean. 2nd (Ed). Department of natural resources. CSIRO publishing. Australia. Pp. 77-79.

Huhtanen, P., Nousiainen, J. and Rinne, M. 2006. Recent developments in forage evaluation with special reference to practical applications. Agric. Food Sci. 15, 293-323.

102

ICIPE. 2017. ICIPE, Technologies and Innovations: Transforming Africa’s agriculture and livelihoods. ICIPE e-bulletin 7 (3-4): 1-14.

IGAD. 2011. The contribution of livestock to the Ethiopian economy-part II, IGAD livestock policy initiative, Roy Behnke, Odessa Centre, Great Wolford, United Kingdom.

Imran, M., Khan, S., Khalid, R., Gurmani, Z.A., , A., Masood, M. & Sultani, M.I. 2007. Performance of different millet cultivars for fodder production under rainfed conditions of Islamabad. Sarhad J. Agric. 23 (2):281-284.

Jank, L., Barrios, S.C., do Valle, C.B., Simeao, R.M. and Alves, G.F. 2014. The value of improved pastures to Brazilian beef production. Crop and Pasture Science, published first online 2014 Mar. 11 at: http://dx.doi.org/10.1071/CP13319.

Kaplan M, Kamalak A, Kasra AA, Güven I. 2014. Effect of maturity stages on potential nutritive value, methane production and condensed tannin content of Sanguisorba minor. Kafkas Univ Vet Fak Der. 20 (3):445-449.

Kassahun Desalegn and Wasihun Hassen.2015. Evaluation of Biomass Yield and Nutritional Value of Different Species of Vetch (Vicia). Jimma University, College of Agriculture and Veterinary Medicine, Department of Animal Science, Jimma, Ethiopia Academic Jthenal of Nutrition 4 (3): 99-105, 2015 ISSN 2309-8902

Khan, Z.R., Midega, C.A., Pittchar, J.O., Murage, A.W., Birkett, M.A., Bruce, T.J. and Pickett, J.A. 2014. Achieving food security for one million sub-Saharan African poor through push–pull innovation by 2020. Philosophical Transactions of the Royal Society B 369(1639): 20120284.

Kizima JB, Mtengeti E J and Nchimbi-Msolla S. 2014. Seed yield and vegetation characteristics of Cenchrus ciliaris as influenced by fertilizer levels, row spacing, and cutting height and season. Livestock Research for Rural Development Volume 26, Article pp148.

Laidlaw A. 2005. The relationship between tiller appearance in spring and contribution of dry- matter yield in perennial ryegrass (Loliumperenne L.) Cultivars differing in heading date. Grass Forage Sci. 2005; 60:200–9.

Landon, J.R. 1991. Booker Tropical Soil Manual. Longman Scientific and Technical Essex, UK. pp.474.

Lardner, H.A., Damiran, D. and McKinnon, J.J. 2015. Evaluation of 3 brome grass species as pasture: Herbage nutritive value, estimated grass dry matter intake and steer performance. Livestock Science, 175, 77–82.

Larry A, Redmon and Mark L. McFarland. 2013. Soil pH and Forage Production.http://soilcrop.tamu.edu .SCS-2013-11. 103

Likawent Yeheyis. 2012. Potential of lupins (Lupinus spp. L.) for human use and livestock feed in Ethiopia. Ph.D. Thesis. Humboldt University of Berlin, Berlin, Germany.

Linn JG, Martin NP. 1999. Forage nutitive value tests and interpretations. University of Minnesota Extension service, Food and Environmental science.

Lonsdale, C. 1989. Raw Materials for Animal Feeds Compounders and Farmers. Chalcombe Publications Great Britain. Pp. 17-47.

Lounglawan, P., W. Lounglawan and W.Suksombat. 2013. Effect of cutting interval and cutting height on yield and chemical composition of King Napier grass (Pennisetumpurpureum x Pennisetumamericanum). APCBEE Procedia, 8:27–31.

Maass BL, Midega AO, Mutimura M, Rahetlah VB, Salgado P, Kabirizi JM, Khan ZR, Ghimire S, Rao IM. 2015. Homecoming of Brachiaria: improved hybrids prove useful for African animal agriculture. East Africa Agric FO J. 2015; 81:71–8

Mathison, G.W., Okine, E.K., Vaage, A.S., Kaske, M. and Millingan, L.P. 1995. Current understanding of the contribution of the propulsive activities in the fore stomach to the flow of digesta. Proceedings of the eighth International Symposium on Ruminant Physiology. Herstellung: Schröders, Agentur, D-14199, Berlin, pp, 23–41.

Maynard, A.L., K.L., Loosli and F.H., Hintz, 1981. Animal Nutrition. Seventh edition. Tata McGraw-Hill Publishing Company Limited. New Delihi, India.

McDonald, P., Edwards, R. A., Greenhalgh, J. D. and Morgan, C. A. 2002. Animal Nutrition 6th edition. United Kingdom: Longman.

McDonald, P., Edwards, R.A., Greenhalgh, J.F.D. and Morgan, C.A. Sinclair, L. A, Wilkinson, R. G., 2010. Animal Nutrition 6th ed. Longman Group UK Ltd, England. 693p.

McDonald, P., Edwards, R.A., Greenhalgh, J.F.D., Morgan, C.A., Sinclair, L.A. and Wilkinson, R.G. 2011. Animal nutrition. Seventh Edition, Pearson Education Limited, Edinburgh Gate, Harlow, Essex CM20 2JE, England, pp692.

McDowelI, L. R. 1976. Mineral deficiencies and toxicities and their effect on beef production.

McDowell LR and Valle G. 2000. Major minerals in forages. In: D.I. Givens, E. Owen, R.F.E. Axford and H. M. Omed (Eds). Forage Evaluation in Ruminant Nutrition. CABI Publishing, New York, NY. pp. 373-397.McGraw Hill International Book Company, London.

Miles JW, do Valle CB, Rao IM, Euclides VPB (2004). Brachiaria grasses. Sollenberger. pp 745–783.

104

Mimila Zemene Mengstu. 2018. Effect of Plant Spacing and Harvesting Age on Plant Characteristics, Yield and Chemical Composition of Brachairia Mutica Grass at Bahir Dar, Ethiopia. MSc. Thesis, Bahir Dar University College of Agriculture and Environmental sciences, pp112.

Minson DJ .1990. Forage in ruminant nutrition. Academic press, Inc. San Diego.

Misganaw Walie, Beyadglign Hunegnaw, Mekonnen Tilahun, Tewodros Bimrew, Wondimeneh Mekonnen, Lisanework Molla, Adebabay Kebede and Asresu Yitayew .2018. Development of Napier Grass Based Feeding Package for Crossbred Dairy Cows at Andassa Livestock Research Center. Proceeding of the 10th Annual Regional Conference on Completed Livestock Research Activities.

Misganaw Walie, Wondmeneh Mekonnen, Beyadglign Hunegnaw, Tekeba Eshetie Agraw Amane, Adebabay Kebede, Simegnew Tamir, and Asresu Yitayew.2016. Participatory cowpea varieties selection and optimum sowing date determination for under sowing on maize crop production in North Western, Ethiopia. Proceeding of the 9th Annual Regional Conference on Completed Livestock Research Activities on 9-20 March, 2015.

Moore, G.; Yates, R.; Barrett-Lennard, P.; Nichols, P.; Wintle, B .2013. Establishment Guide for Sub-Tropical Grasses—Key Steps to Success; DAFWA: Kensington, Australia, 2013.

Moreta D.E., Arango J., Sotelo M., Vergara D., Rincón A., Ishitani N., Castro A., Miles J., Peters M., Tohme J., Subbarao G.V., Rao I.M. 2014. Biological nitrification inhibition (BNI) in Brachiaria pastures: A novel strategy to improve eco-efficiency of crop- livestock systems and to mitigate climate change.Tropical Grasslands–Forrajes Tropicales, 2: 88–91.

Mulaa, M.; Kabirizi, J.; Pallangyo, B.; Hanson, J.; Proud, J.; Mukiibi, E.; Maeda, C.; Wanjala, B.; Awalla, B.J and Namazzi, C. 2013.Diversity, biomass and resistance to stunt in Napier grass clones in East and Central Africa region. Association for Strengthening Agricultural Research in Eastern and Central Africa, Entebbe, Ugandapp2734.ISBN:978-92-95070-96-7

Mupenzi Mutimura, Cyprian Ebong, Idupulapati M. Raoand Ignatius V. Nsahlai .2017. Effect of cutting time on agronomic and nutritional characteristics of nine commercial cultivars of Brachiaria grass compared with Napier grass during establishment under semi-arid conditions in Rwanda.Afr. J. Agric. Res, Vol. 12(35), pp. 2692-2703, 31 August, 2017.

Mustaring, I. Subagyo, Soebarinoto and Marsetyo .2014. Growth, yield and nutritive value of new introduced Brachiaria species and legume herbs as ruminant feed in central Sulawesi, Indonesia .Pakistan j. Agric. Res. Vol. 27 no.2, 2014.

105

Mutimura .M and Everson. T .2012. On-farm evaluation of improved Brachiaria grasses in low rainfall and aluminium toxicity prone areas of Rwanda.International Journal of Biodiversity and Conservation Vol. 4(3), pp. 137-154, March 2012.

Nafatul Umami, Meita Puspa Dewi , Bambang Suhartanto, Cuk Tri Noviandi, Nilo Suseno, Genki Ishigaki, Ryo Akashi .2015. Growth and Productivity of Brachiaria brizantha CV MG 5 under the effect of different dose of NPK fertilizer. The 6th International Seminar on Tropical Animal Production Integrated Approach in Developing Sustainable Tropical Animal Production October 20-22, 2015, Yogyakarta, Indonesia.

Negasu Yibarkew Getu .2018. Effects of fertilizer type and plant spacing on Plant Characteristics, Yield and Chemical Composition of Brachiaria hybrid mulato II grass grown in the lowland of northwest Ethiopia . MSc. Thesis, Bahir Dar University College of Agriculture and Environmental sciences.

Nguku Susan. 2015. An evaluation of brachiaria grass cultivars productivity in semi-arid Kenya. MSc. Thesis, South Eastern Kenya University, Kenya.

Njarui, D. M.G., Gichangi, E. M., Gatheru, M., Njunie, M.N., Ondiko, C. N., Nyambati, E. M., Ndungu K., Muinga R. W., and Mureithi, J. G. 2015. Climate smart Brachiaria for improving livestock production in E. Africa- KALRO. Progress report presented in mid- year review meeting held on 2-3 July 2015 at Rubona, Rwanda. pp. 361-367.

Nkongolo KK, Chinthu KKL, Malusi M, Vokhiwa Z .2008. Participatory variety selection and characterization of sorghum (Sorghum bicolor (L.) Moench) elite accessions from Malawian gene pool using farmer and breeder knowledge. Afric. J. Agricult. Res., 3(4): 273–283.

NRC. National research council. 2001. Nutrient Requirement of Domestic Animals, No.4: Nutrient Requirement of Dairy Cattle, seventh rev: Ed, National Academy Press, Washington Dc, USA.

Okalebo, J. R., Gathua, K. W., Woomer, P. L. 2002. Laboratory Methods of Soil and Plant Analysis: A Working Manual. TSBF-KARI, SSEA, SACRED Africa, Nairobi, Kenya, 2002 pp. 128.

Olsen, S.R., Cole, C.V., Watenabe, F.S. and Dean, L.A. 1954. Estimation of the valuable phosphorus in soil by extraction with sodium bicarbonate United States Department of Agriculture Circ., 939.USA

Omer Suha Uslu, Ozer Kurt, Emrah Kaya and Adem Kamalak. 2018. Effect of species on chemical composition, metabolizable energy, organic matter digestibility and methane production of some legume plants grown in Turkey.Journal of Applied Animal Research.2018, Vol. 46, No. 1.

106

Patrick C. Hoffman. 2005. Ash Content of Forages.UW-Madison Dairy Science Department Marshfield Agricultural Research. Focus on Forage - Vol 7: No. 1.

Peiretti PG; Gai F; Tassone S. 2015.Nutritional value and fatty acid profile of Niger (Guizotia abyssinica) plant during its growth cycle. Livestock Research for Rural Development 27, article pp18. www.lrrd.org/lrrd27/1/ peir27018.htm.

Pizarro, E.A, M. Hare, J.H. Antezana Rojas, R.R. Ramón, I.G. Miranda, A. Chávez Chena, A. Balbuena, and J.W. Mile. 2008. Harvest methods and seed yield potential in Brachiaria hybrids.

Raanjhman SK. 2001. Animal nutrition in the tropics. 5th (Ed). New Delhi: Vikas Publishing House; p. 593.

Ranjhan SK. 1993. Animal Nutrition in the Tropics. (3rdedn). Vikas Publishing House Pvt. Ltd., New Delhi.

Rao I., Ishitani M., Miles J., Peters M., Tohme J., Arango J., Moreta D.E., Lopez H., Castro A., Hoek R.V.D., Martens S., Hyman J., Tapasco J., Duitam J., Suárez H., Borrero G., Núñez J., Hartmann K., Domínguez M., Sotelo M., Vergara D., Lavelle P., Subbarao G.V., Rincon A., Plazas C., Cadisch G.,Mendoza R., Rathjen, L., Karwat H. 2014. Climate-smart crop-livestock systems for smallholders in the tropics:. Tropical Grasslands–Forrajes Tropicales, 2: 130–132.

Rasby, R.J., Berger, A.L., Bauer, D.E. and Brink, D.R. 2011. Minerals and vitamins for beef cows. Know how, Extension materials, University of Nebraska, pp7.

Rayburn, E. B., Lozier, J. D., Sanderson, M. A., Smith, B. D., Shockey, W. L., Seymore, D. A.,and Fultz, S. W. (2007). Estimating Pasture Forage Mass from Pasture Height. Extension Service West Virginia University, 1, 1-2.

Riet-Correa, F. 2007. Intoxicações por plantas e micotoxinas. Doenças de Ruminantes eEqüídeos 2007, 2, 99–219.

Rodrigues, R.C., Sousa, T.V.R., Melo, M.A.A, Araújo, J.S., Lana, R.P., Costa, C.S., Oliveira, M.O., Parente, M.O.M. and Sampaio, I.B.M. 2014. Agronomic, morphogenic and structural characteristics of tropical forage grasses in northeast Brazil. Tropical Grasslands – Forrajes Tropicales. 2: 214-222.

Santos, M.V.F., J.C.B. Dubeux Jr., M.C. Silva, S.F. Santos, L.C. Ferreira, A.C.L. Mello, I. Farias, E.V. Freitas. 2003. Productivity and chemical composition of tropical grasses in the zone of Pernambuco. (In Portuguese, with English Abstract.) Revista Brasileira de Zootecnia, Vol.32: 821-827.

SAS® institute inc. 2002. SAS user’s guide: statistics, SAS inst. Inc., Cary, NC, USA.

107

Saurav Dwari and Amal Kumar Mondal. 2011. Systematic studies (morphology, anatomy and palynology) of economically viable grass Brachiaria Mutica (Forsskil) Stapf in Eastern India. African Journal of Plant Science Vol. 5 (5), pp. 296-304, May 2011.

Schroeder, J. 2012. Interpreting Forage Analysis. North Dakota: NDSU Extension Service

Seyoum B, Getinet A, Abate T, Dereje F, 2001. Present status and future direction of feed resources and nutrition research targeted for wheat based crop livestock production system in Ethiopia. In: wall pc (Ed) wheat and weed: food and feed. Proceedings of two- stakeholder workshop on improving the productivity of crop livestock production in wheat based farming system in Ethiopia, Addis Ababa, Ethiopia, Pp 207-226.

Shigdaf Mekuriaw Beyadglign Hunegnaw, Agraw Aamane, Lisanework Molla, Asresu Yitayew and Likawent Yeheyis. 2018. Tree Lucerne Leaves Substitution of Noug Seed Cake as Protein Supplementation on Growth Performance of Washera Sheep in Awi Zone, Amhara Region, Ethiopia. Proceeding of the 10th Annual Regional Conference on Completed Livestock Research Activities. Pp 178-194.

Shimelis Mengistu and Temesgen Alene. 2016. Training report on forage production, feed management and utilization for Africa RISING project farmers in Basona Worena, Ethiopia. International Livestock Research Institute (ILRI).

Shimelis Mengistu, 2018. Assessment Of Forage Production, Feed Resource Utilization And Substitution Effect Of Oat-Vetch Forage For Concentrate Mix On Performance Of Sheep Fed Desho Grass As A Basal Diet In Damot Gale District Of Wolaita Zone, SNNPR. A Thesis Submitted to the School Of Animal and Range Scineces, College Of Agriculture, School Of Graduate Studies, Hawassa University.

Solomon Bogale, Solomon Melaku and Alemu Yami, 2009. The interdependence of crop- livestock production sectors: the case of Sinana-dinsho district in Bale highlands of Ethiopia. Tropical and Subtropical Agroecosystems. Vol 42(2).

Smart AJ; Schacht WH; Moser LE; Volesky JD. 2004. Prediction of leaf/stem ratio using near- infrared reflectance spectroscopy (NIRS): A Technical Note. In: Agronomy & Horticulture Faculty Publications, Vol. 39. http://goo.gl/ QFvzF9.

Steel, R. D. G. and J. H. Torrie. 1986. Principles and Procedures of Statistics. A Biometrical Approach. New-York: 5th Ed. McGraw-Hill International Book Company.

Subbarao G.V., Nakahara K., Hurtado M.P., Ono H., Moreta D.E., Salcedo A.F., Yoshihashi A.T., Ishikawa T., Ishitani M., Ohnishi-Kameyama M., Yoshida M., Rondon M., Rao I.M., Lascano C.E., Berry W.L., Ito O. 2009. Evidence for biological nitrification inhibition in Brachiaria pastures. Proc. Natl. Acad. Sci. USA, 106: 17302-17307.

Sultan, J. I., I. Rahim, H. Nawaz and M. Yaqoob. 2007. Nutritive value of marginal land grasses of northern grassland of Pakistan. Pakistan J. Bot.39, 1071-1082. 108

Susan A. Nguku, Donald N. Njarui, Nashon K.R. Musimba, Dorothy Amwata and Eric M.Kaindi. 2016. Primary Production Variables of Brachiaria Grass Cultivars in Kenya Dry lands. Tropical and Subtropical Agro ecosystems, 19 (2016): 29 – 39.

Tavirimirwa B., Manzungu E. and Ncube S. 2012. The evaluation of dry season nutritive value of dominant and improved grasses in fallows in Chivi district, Zimbabwe. Online J. Anim. Feed Res.2 (6).http://www. Science-line. Com/index/; http://www. ojafr.ir, 470- 474.

Teixeira, F.A.; Bonomo, P.; Pires, A.J.V.; da Silva, F.F.; Fries, D.D.; da Hora, D.S. 2011. Produção anual e qualidade de pastagem de Brachiaria Decumbens diferida e estratégias de adubação nitrogenada. Anim. Sci. 2011, 33, 241–248.

Tekalign Tadesse. 1991. Working Document: Soil, Plant, Water Fertilizer, Animal Manure and Compost Analysis Manual, International Livestock center for Africa, No. B13, Addis Ababa, Ethiopia.

Tessema Zewudu, 2008. Effect of plant density on morphological characteristics, yield and chemical composition of Napier grass (Pennisetum purpureum (l.) Schumach). East African journal of sciences, volume 2 (1) 55-61.achievements in northwestern Ethiopia. Pp. 62-69.

Tessema Zewudu, Ashagre A and Solomon Mengistu. 2010. 'Botanical composition, yield and nutritional nutitive value of grass land in relation to stages of harvesting and fertilizer application in the highlands of Ethiopia. African journal of range and forage science 27 (3): 117 -124

Tessema Zewdu, Baars, R. and Alemu Yami, 2002. Effect of plant height at cutting, source and level of fertilizer on yield and nutritional nutitive value of Napier grass (Pennisetum purpureum (L.) Schumach.). African Journal of Range Science. 19: 123-128

Tessema Zewudu, Bears R.M.T. and Alemu Yami. 2003. Effect of plant height at cutting and fertilizer on growth of Napier Grass (Pennisetum purpureum). Tropical Science, 42, 57- 61.

Tessema, Z. K., Boer, W. F. de, Baars, R. M., and Prins T. 2011. Changes in soil nutrients, vegetation structure and herbaceous biomass in response to grazing in a semi-arid savanna of Ethiopia. J Arid Environ 75: 662-670.

Thornton, P. K. and Herrero, M. 2010. Potential for reduced methane and carbon dioxide emissions from livestock and pasture management in the tropics. Proceedings of the National Academy of Sciences, 107(46), 19667-19672.

Tilley, J. M. A and Terry ,R. A. ( 1963), A two-stage technique for the in vitro dry matter digestibility of forage crops. Journal of British Grassland Society, 18: 104-111. 109

Tudsri S, Jorgensen ST, Riddach P, Pookpakdi A (2002). Effect of cutting height and dry season closing date on yield and nutitive value of five Napier grass cultivars in Thailand. Trop. Grassland, 36: 248–252.

Ulfina G, Habtamu A, Jiregna D and Chala M, 2013. Utilization of brewer’s waste as replacement for maize in the ration of calves. Http://www.researchwebpub.org/wjar.

Urio, N.A., S.V. Sarwattand E.J. Mtengeti .2006. A review of potential of Brachiaria species in Tanzania.

Van Saun RJ. 2006. Determining forage nutitive value: understanding feed analysis. Lamalink.com. 2006; 3(8):18–9.

Van soest, P. J., H. B. Robertson and B. A. Lewis, 1991. Method of dietary fiber and non-starch polysaccharides in relation to animal material. J. Dairy sci. 74:3583-3587

Van Soest. P. J. 1994. Nutritional Ecology of Ruminants. 2nd Edition. pp. 50-65, 478. New York, USA: Cornel University Press: Ithaca Pp. 50-65, 478.

Vendramini, J.M.B., A.T. Adesogan, M.L.A. Silveira, L.E. Sollenberger, O.C. Queiroz, and W.E. Anderson. 2010. Nutritive value and fermentation parameters of warmseason. Prof. Anim. Sci. 26, 193-200.

Vendramini J., Sellers, B., Sollenberger, L.E., and Silveira, M. 2011. Mulato II (Brachiaria sp.) SS AGR 303. Florida, USA: UF/IFAS Extension.

Vendramini JMB, Sollenberger LE, Soares AB, Da Silva WL, Sanchez JMD, Valente AL, Aguiar AD, Mullenix MK .2014. Harvest frequency affects herbage accumulation and nutritive value of Brachiaria grass hybrids in Florida. Trop. Grasslands – Forrajes Trop. 2:197-206.

Vilela, L., G.B. Martha Jr., L.G. Barioni, A.O. Barcellos. 2004. Fertilizer in the recovery and intensification of animal production on pastures. In: Symposium on management of pastures, 21. (In Portuguese, with English Abstract.) Piracicaba, Anais. FEALQ. p. 425- 472.

Walkley, A.J. and Black, I.A. 1934. Estimation of soil organic carbon by the chromic acid titration method. Soil science, 37:29-38

Wasson. P.A, R.A. Richards1, R. Chatrath2, S.C. Misra3, S.V. Sai Prasad, G.J. Rebetzke1, J.A. Kirkegaard1, J. Christopher5 and M. Watt. 2012. Traits and selection strategies to improve root systems and water uptake in water-limited wheat crops. Journal of Experimental Botany, Page 1 of 14 doi:10.1093/jxb/ers111

Wondimeneh Mekonnen, Likawent Yeheyis, Beyadglign Hunegnaw, Meseganaw Walle,Tekaba Eshetie , Simegnew Tamir, Agraw Amane and Mekonnen Tilahun . 2016. 110

Participatory Variety Selection of Different Sweet Lupin (Lupinus angustifolius L.)Cultivars for Under Sowing on Maize (Zea mays L.) Crop Production in North West, Ethiopia. Proceeding of the 9th Annual Regional Conference on Completed Livestock Research Activities.

Worku B., Denbela H., T/yohanis B. 2017. Effect of planting space and fertilizer rate on productivity of Desho grass (Pennisetum pedicellatum) in Jinka agricultural Research center, Southern Ethiopia. American Journal of Agricultural Research (ISSN: 2475- 2002.

Wubetie Adnew Wassie, Berhanu Abraha Tsegay, Asaminew Tassew Wolde and Bimrew Asmare Limeneh. 2018b. Evaluation of morphological characteristics, yield and nutritive value of Brachiaria grass ecotypes in northwestern Ethiopia. Agric & Food Secur (2018) 7:89.

Wubetie Adnew, Berhanu Tsegay, Asaminew Tassew and Bimrew Asmare. 2018a. Assessments of farmers′ perception and utilization status of Brachiaria grass in selected areas of Ethiopia. BIODIVERSITAS ISSN: 1412-033X. Pages: 955-966 DOI: 10.13057/biodiv/d190326.

Yasin, M., Malik, M.A., and Nazir, M.S. 2003. Effect of different spatial arrangement on forage yield, yield components and nutitive value of Mott Elephant grass. Pakistan J. agro. 2(1): 52-58.

Yayneshet Tesfaye. 2010. Feed resources available in Tigry region, north Ethiopia, for production of 18. Export nutritive value meat and livestock Addis Ababa Ethiopia, Pp: 9-16.

Yoseph Mekasha, Mengistu Urge, Mohammed Y. Kurtu and MergaBayissa. 2010. Effect of harvesting stage on yield and quality of natural pasture in the central high lands of Ethiopia. In: Proceedings of 3rd national conference of the Ethiopian society of animal production.27-29, April 1995.pp.316-322. Addis Ababa, Ethiopia.

Yusuf AO, Muritala RO. 2013. Nutritional evaluation and phytochemical screening of common plants used in smallholder farming system. Pac J Sci Technol. 14(2):456–462.

111

Chapter 7. APPENDICES

7.1. Summary on Analysis of Variance of all experimental cultivars

Appendix 7.1. 1 Summary on Analysis of Variance for morphology and dry matter yield of Brachiaria grass cultivars

Parameters DF Sum Mean F Value Pr > F SL Root MSE Coeff Var R- Squares Square Square PH 19 7.402 0.390 318.03 <.0001 *** 0.035 5.6 0.993 NTPP 19 2544.349 133.913 80.97 <.0001 *** 1.286 3.747 0.976 NLPP 19 314648.22 16560.43 54.46 <.0001 *** 17.44 8.196 0.963 LLPP 19 593.331 31.228 121.22 <.0001 *** 0.508 3.176 0.983 NRPP 19 11926.802 627.726 42.08 <.0001 *** 3.862 2.788 0.952 LRPP 19 429.926 22.628 38.68 <.0001 *** 0.765 7.154 0.948 LSR 19 7.663 0.403 8.71 <.0001 *** 0.215 13.24 0.805 DMY 19 1535.171 80.798 790.44 <.0001 *** 0.32 3.779 0.997

PH= Plant height, NTPP= number of tiller per plant, NLPP=number of leaves per plant, LLPP= leaf length per plant, NRPP= number of roots per plant, LRPP= length of root per plant, LSR= leaf to stem ratio, DMY= dry matter yield

112

Appendix 7.1. 2 Summary on Analysis of Variance for chemical composition and digestibility of Brachiaria grass cultivars

Parameters DF Sum Mean Square F Value Pr > F SL Root MSE Coeff Var R - Squares Square

DM (%) 19 3.329 0.175 15.63 <.0001 *** 0.106 0.113 0.937 Ash (%) 19 104.404 5.495 76.61 <.0001 *** 0.268 2.306 0.986 OM (%) 19 104.404 5.495 76.61 <.0001 *** 0.268 0.303 0.986 NDF (%) 19 823.821 43.359 43.63 <.0001 *** 0.997 1.421 0.976 ADF (%) 19 907.487 47.762 77.22 <.0001 *** 0.786 1.499 0.986 ADL (%) 19 31.199 1.642 37.08 <.0001 *** 0.210 2.097 0.972 IVDM (%) 19 1545.513 81.343 34.03 <.0001 *** 1.546 2.548 0.969 DOMD (%) 19 1179.228 62.065 24.76 <.0001 *** 1.583 3.003 0.959 CP (%) 19 175.74 9.249 315.52 <.0001 *** 0.17 1.62 0.996 CPY (t/ha) 19 16.3 0.85 1408 <.0001 *** 0.024 2.58 0.999 ME (MJ/kg) 19 17.603 0.926 21 <.0001 *** 0.210 2.721 0.952

DM= Dry matter content, ASH= ash, OM= organic matter, NDF=Neutral detergent fiber, ADF=Acid detergent fiber, ADL=Acid detergent lignin, CP= Crude protein, CPY= Crude protein yield, IVDMD= In vitro dry matter digestibility, IVOMD= In vitro organic matter digestibility, ME = Metabolizable energy

113

Appendix 7.1. 3 Summary ANOVA on main and interaction effects on morphology and DMY of Brachiaria grass

Variables Parameters PH NTPP NLPP LLPP NRPP LRPP LSR DMY t/ha Cultivars (C) *** *** *** *** *** *** *** *** Mutica 1.24a 33.56bc 226.51b 17.93a 121.682c 10.35b 1.07d 12.45a Mulato-II 0.65b 40.59a 274.1a 15.34c 145.25a 12.625a 1.844a 11.65b Mulato-I 0.446c 34.38b 211.04b 14.68d 143.33a 9.962b 1.735b 8.4c Marandu 0.43c 30.28d 160.14d 14.87cd 14.83a 10.325b 1.92a 5.356d La Liberated 0.402c 32.79c 192.03bc 17.05b 137.54b 10.467b 1.5624c 4.45e Overall mean 0.624 34.321 212.759 15.981 138.528 10.692 0.0487 8.459 SEM 0.042 0.863 9.19 0.409 3.423 0.3387 13.23 0.6314 Cv 5.601 3.74 8.19 3.175 2.788 7.153 0.805 3.779 Soil type (S) *** *** *** *** *** *** Ns *** Black 0.66a 35.69a 218.6a 16.64a 137.65b 11.42a 1.603 8.824a Red 0.59b 32.95b 206.92b 15.34b 139.6a 9.96b 1.651 8.094b Fertilizer (F) *** *** *** *** *** *** *** *** With 0.75a 37.8a 253.37a 17.31a 148.37a 12.91a 1.771a 11.679a Without 0.5b 30.84b 172.14b 14.65b 128.69b 8.47b 1.481b 5.239b Interaction *** *** S*F 0.5294ns p<.0001 p<.0001*** 0.4468ns 0.0046** 0.1169ns 0.7936ns p<.0001 S*C p<.0001** p<.0001 0.0929ns p<.0001*** p<.0001*** 0.0051** 0.7124ns p<.0001 F*C p<.0001** p<.0001 p<.0001*** p<.0001*** 0.0001*** p<.0001*** 0.005* p<.0001 C*S*F 0.001** p<.0001 p<.0001*** p<.0001*** 0.0108* 0.0066** 0.0251* p<.0001

PH= Plant height, NTPP= number of tiller per plant, NLPP=number of leaves per plant, LLPP= leaf length per plant, NRPP= number of roots per plant, LRPP= length of root per plant, LSR= leaf to stem ratio, DMY= dry matter yield, S*F interaction of soil type and fertilizer, S*C= interaction of soil and cultivars, F*C= interaction of fertilizer and cultivars , C*S*F= interaction of cultivars, soil type and fertlilizer

114

Appendix 7.1. 4 Summary ANOVA on main and interaction effects on chemical composition and digestibility of Brachiaria grass

Variables Parameters DM ASH OM CP NDF ADF ADL DMD OMD ME CPY Cultivars (C) *** *** *** *** *** *** *** *** *** *** *** Mutica 93.8a 9.2c 89a 9.1 77.4a 65.54a 11.3a 51.7c 45.3c 6.9c 1.22b Mulato-II 93.4b 11.b 88.8b 10.97b 68,7c 50.5b 9.8b 60.6b 52.5b 7.73b 1.45a Mulato-I 93.73a 11.2b 86.08c 10.8b 65.8d 49d 9.9b 68.2a 58.7a 8.7a 0.96c Marandu 93.4b 13.02a 86.78c 12.14a 68,6c 50.5b 9.8b 61.5b 53.3b 7.68b 0.67d La Liberated 93.41b 13.12a 86.8c 9.7c 70.54b 51b 9.3c 61.6b 52.9b 7.6b 0.45e Overall mean 93.55 11.6 88.4 10.55 70.13 52.3 10.04 60.7 53.3 7.73 0.96 Cv 0.13 2.3 0.36 1.6 1.4 1.5 2.1 2.55 3 2.7 2.5 Soil type (S) *** *** *** *** *** *** *** *** *** *** *** Black 93.5b 11.5b 88.5a 10.5b 70.5a 53.3a 10.2a 59.8b 52.1b 7.7b 1.02a Red 93.6a 11.7a 88.2b 10.6a 69.7b 51.7b 9.9b 65.6a 53.4a 7.8a 0.99b Fertilizer (F) *** *** *** *** *** *** *** *** *** *** *** With 93.4b 11.75a 88.25b 11.95a 69.5b 52.3b 9.9b 62.2a 54.2a 7.9a 1.36a Without 93.6a 11.5b 88.5a 9.15b 70.7a 52.7a 10.2a 59.2b 51.2b 7.5b 0.6b Interaction *** *** *** *** *** *** *** *** *** *** *** S*F p<.0001 p<.0001 p<.000 0.0002 p<.0001 p<.0001 p<.0001 0.00014 <.0001 0.001 <.000 1 7 1 S*C 0.0002 p<.0002 p<.000 p<.0001 p<.0001 p<.0001 p<.0001 0.0001 <.0001 <.000 <.000 1 1 1 F*C 0.0031 p<.0001 p<.000 p<.0002 p<.0001 p<.0001 p<.0001 p<.0001 <.0001 <.000 0.004 2 1 C*S*F p<.0001 0.0001 p<.000 p<.0001 p<.0001 p<.0001 p<.0001 p<.0001 0.004 0.009 <.000 1 1 DM= Dry matter content, ASH= ash, OM= organic matter, NDF=Neutral detergent fiber, ADF=Acid detergent fiber, ADL=Acid detergent lignin, CP= Crude protein, CPY= Crude protein yield, IVDMD= In vitro dry matter digestibility, IVOMD= In vitro organic matter digestibility, ME = Metabolizable energy, S*F interaction of soil type and fertilizer, S*C= interaction of soil and cultivars, F*C= interaction of fertilizer and cultivars , C*S*F= interaction of cultivars, soil type and fertlilizer

115

Appendix 7.1. 5 Pair-wise ranking of Brachiaria grass cultivars by participant farmers’ at black soil (N= 20)

Mutica Mulato- II Mulato- I Maranda La Liberated to 2 2 2 to 2 Criteria 3 4 5 3 4 5 To 1 3r 4 5 tot 1 3 4 5t 3 4 To 1st 2nd ta 1st n n n ta 1st n 5th rd th th rd th th tal st d th th al st rd th h rd th tal l d d d l d 1 Plot 1 12 1 12 6 4 16 0 5 0 2 4 0 4 0 0 0 6 0 0 0 2 5 6 0 0 0 0 14 70 cover 6 0 8 0 5 5 Tiller 7 2 2 1 6 0 18 0 0 0 0 0 0 0 80 0 2 0 0 88 0 0 0 0 0 0 0 4 10 56 number 2 0 0 6 4 Plant 1 6 1 1 1 1 5 10 0 0 0 0 0 0 60 0 0 5 0 69 0 0 2 0 0 0 0 0 5 15 height 0 0 0 0 8 5 1 3 1 1 1 1 1 6 Leafiness 0 0 5 5 5 0 0 0 56 2 0 0 0 24 0 0 0 0 0 0 0 15 30 0 8 0 0 5 5 0 Smoothn 1 1 2 1 1 2 1 0 0 0 0 5 0 0 0 25 0 5 5 0 25 0 0 5 0 0 0 0 0 10 es 5 5 0 5 5 0 0

Total 3 2 score 0 34 32 6 18 2 1 6 0 1 Rank 3 1 2 4 5

116

Appendix 7.1. 6 Pair-wise ranking of Brachiaria grass cultivars by participant farmers’ at red soil (N=16)

Mutica Mulato –II Mulato -I Maranda La Liberated to Criteria 2 3 4 5 1s 2 3 4 5 To 1 2 3 4 5 To 1 2 3 4 5 To 1 2 3 4 5 To 1st ta nd rd th th t nd rd th th tal st nd rd th th tal st nd rd th th tal st nd rd th th tal l Tiller 1 7 1 11 1 1 2 0 0 0 4 4 0 0 0 0 5 0 85 0 0 0 4 75 0 0 0 5 4 45 number 2 0 4 0 2 1 Leafines 1 5 1 1 10 0 0 0 0 0 0 0 0 2 6 7 0 60 0 0 0 5 7 48 0 0 0 4 9 52 s 4 6 6 0 4 Plant 4 1 1 8 8 0 0 0 8 8 0 0 0 48 0 0 6 0 60 0 0 2 0 36 0 0 0 0 6 18 height 8 4 0 plot 2 1 1 1 0 0 4 4 4 8 0 0 0 44 2 8 6 0 52 0 0 2 6 0 16 0 0 0 0 24 cover 4 4 0 2 Smooth 1 1 1 1 1 0 0 0 4 0 0 0 0 16 0 0 0 0 16 0 0 4 2 22 0 0 0 8 2 10 nes 2 6 6 6 6 2 total 1 32 27 19 14 score 4 2 3 7 9 rank 3 1 2 4 5

117

Appendix 7.1. 7 Summary of on survival rate of experimental Brachiaria grass cultivars

cultivars Black soil (Andassa) Red soil (Mecha) FER N-planted N-survived Survival rate N-planted N-survived Survival rate (%) B.mutica WF 30 30 100 30 30 100 B.Mulato-II WF 30 29 96.67 30 27.83 92.78 B .Mulato-I WF 30 27.8 92.78 30 26.5 88.33 B .Marunda WF 30 25.67 85.56 30 26.17 87.22 La Liberated WF 30 26.5 88.3 30 25.17 83.9 B .mutica WO 30 30 100 30 30a 100 B .Mulato-II WO 30 28.33 94.44 30 26.5 88.33 B .Mulato-I WO 30 27d 90d 30 25.3 84.33 B .Marundu WO 30 26.17 87.2 30 24.5 81.67 B .La Liberated WO 30 23.77 79.22 30 24.17 80.56 B.decumbens WF 30 3 10 30 Ne Ne B.decumbens WO 30 2i 6i 30 Ne Ne

Ne= not established, WF= with fertilizer, WO= without fertilizer, N-= numbe

118

7.2. List of Figures

800.0

700.0

600.0

500.0

400.0

300.0

200.0

100.0 Rain fall in mm in fall Rain

0.0 JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC

Months

Andassa Mecha

Figure 7.2. 1 Monthly rainfall data of experimental districts

35.0

30.0

25.0 c 20.0

15.0

10.0

5.0 Tempreture in 0 in Tempreture 0.0 JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC

Months Andassa max Mecha max Andassa min Mecha min

Figure 7.2. 2 Maximum and minimum temperature of experimental districts

119

Figure 7.2. 3 Pictures of experimental Brachiaria cultivars

Figure 7.2.4 farmers perception on experimental cultivars

120