Reducing Booster-Pump-Induced Contaminant Intrusion in Indian Water Systems with a Self-Actuated, Back-Pressure Regulating Valve by David Donald James Taylor BASc., University of Toronto (2011) Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2014 c Massachusetts Institute of Technology 2014. All rights reserved. β Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Department of Mechanical Engineering May 9, 2014 Certified by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alexander H. Slocum Pappalardo Professor of Mechanical Engineering Thesis Supervisor Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . David E. Hardt Chairman, Department Committee on Graduate Theses 2 Reducing Booster-Pump-Induced Contaminant Intrusion in Indian Water Systems with a Self-Actuated, Back-Pressure Regulating Valve by David Donald James Taylor Submitted to the Department of Mechanical Engineering on May 9, 2014, in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Abstract Intermittently-operated water systems struggle to equitably and effectively distribute clean water to customers. One common customer response to intermittency is to supplement the water system’s pressure by using a household, or residential, booster pump. When such booster pumps are directly connected to the water utility’s supply pipe, without an underground isolation tank (sump), they often induce negative pressure in the supply pipe which increases the flow rate. Unfortunately, where leakage rates are high, this negative pressure also increases the risk of contaminant intrusion. This thesis presents the iterative design and field testing of a patent-pending, full-bore, back-pressure regulating valve. The valve’s simple mechanism relies on a stabilized collapsing tube, or ‘Starling Resistor,’ which when installed at a customer’s connection, controls the flow rate and prevents booster pumps from creating negative pressure in the supply pipe. In collaboration with the Delhi Jal Board and several private partners, the valve’s performance was verified in two rounds of field trials in neighborhoods of New Delhi, India including Pitampura, Azad Market, Vivek Vihar, Malvia Nagar, and Vasant Vihar. Using a crossover study, the valve was found to reduce the total contamination risk across all 19 tested houses during supply times by a median of 80%. The valve prevented 96% of pressure below -1 meter and an average of 53 minutes per day, per connection of total negative pressure. In an estimated worst-case scenario for contaminant intrusion, the presence of the valve reduced the contamination risk by two orders of magnitude at six customer connections — enough to correspond to significant reductions in health risks. Thesis Supervisor: Alexander H. Slocum Title: Pappalardo Professor of Mechanical Engineering 3 4 Acknowledgments [Noah’s] Ark should float in Grade 1, and by the time students leave Grade 12..., someone has to have sunk the ark for them! (Denis Lamoureux) I am eternally grateful to both my parents for the blessing of both floating and sinking Noah’s Ark, simultaneously and perpetually. You have role modeled for me that intellectual rigor is the beginning of faith, not its end. You have given me a sticky and wrestling faith; thank you. To my sister for promising to fly down to Boston and slap me in the face if I became pretentious, to my brother for sharing your love and humour with me always, and to both of you, my best friends: front seats should always be designed for three. To my former colleagues at HydraTek & Associates, thank you for teaching me so much of what I know about water and water systems; you allowed me to approach this problem with the confidence of familiarity. To Professor Gutowski, thank you for allowing me to defer my admission to MIT, my thesis would not have been what it is without the experiences I was able to have before coming here. To Professor Slocum, I am grateful for your wise suggestions, great design advice, but most of all for your fierce care and compassion for your students. To all my fellow graduate students in PERG, I often tell people my thesis and design are like the story of Stone Soup: you were frequent contributers of ideas and suggestions to this soup; it could not have happened without you. For all of my lunch-table friendships, I am especially grateful to Anthony Wong for persistently and warmly insisting that “everyone needs to eat; come have lunch.” I very grateful to the MIT Tata Center and to the Tata Trust for the funding, freedom, and fellowship that enabled this research. To Dr. Nevan Hanumara, thanks for caring deeply about my project, always striving to connect me to the right person, and carefully editing many documents for me. To Mark Belanger and the Edgerton Student Shop, you enabled ideas to take form 5 and be tested; thank you Mark, for your humour, warmth, and great help. Thank you to the four Sloan Business students who helped develop a commercialization plan and facilitated many great connections in Delhi: Michael Alley, Ross Frasier, Deepika Goyal, and Horacio Leal. Additional thanks go to Sally Miller and Wesley Cox who helped assemble the alpha prototypes. To my supporters in India, I owe a special set of thank-yous, for as a foreigner I would neither have identified this project, nor been able to test my product without your help and gracious hospitality. This topic and project would not have seen the light of day without the help, advice, and guidance of Mr. Manish Kumar, of Tata Consulting Engineers. I am grateful to you and the whole TCE Delhi office for sharing their your, perspective and expertise with me. I am indebted to Mr. Babbar, from the Delhi Jal Board (DJB), for your patience, wisdom, and support for this project from its inception. To the DJB’s Pitampura staff, you were instrumental in gathering the January field data. Specifically I am grateful to Mr. Thakur, Mr. Sandeep Sharma, and Mr. Yogendra Singh for not only connecting me to Mr. PK Jain, but for also allowing two valves to be installed on your office connections. I am additionally grateful to Mr. PK Jain and especially to Mr. Tej Pal who facilitated nine household connections inside the DJB colony. I owe an enormous thanks to each and every household that gave their informed consent to the water utility, without your permission, this thesis would not have been possible. Specifically, I am grateful to Mr. Ankush, for not only allowing me to test the valve on your house, but for introducing me to all your neighbors and making all five Azad Market connections possible. Further, I am grateful to Mr. Manoj, your willingness to innovate and try new products is contagious — your East Delhi house continues to have a working alpha prototype connected to it; thank you. To the supporters of the March field trial in South Delhi, I am grateful for the partnership and field assistance, specifically Mr. Mappa and especially to Mr. Kévin Beulé and Mr. Ashish Shenoy for your patient, enduring, and innovative help. Additionally I am grateful to Mr. El Hassane and to Mr. Saurabh, for your insights on 6 the project and assistance gathering data. 7 THIS PAGE INTENTIONALLY LEFT BLANK 8 Contents 1 Introduction 23 2 Background 25 2.1 Contaminant intrusion 2.3 2.4 3 4 25 2.1.1 Leakage rates indicate the pathway size for contaminant intrusion 26 2.1.2 Contaminants may be present because of the unplanned locations of water pipes . . . . . . . . . . . . . . . . . . . . . . . . 28 Pipe pressure can deter contamination . . . . . . . . . . . . . 28 Intermittent water systems . . . . . . . . . . . . . . . . . . . . . . . . 30 2.2.1 . . . . . . . . . . . . . . 31 Residential booster pumps . . . . . . . . . . . . . . . . . . . . . . . . 33 2.3.1 Booster pump configurations . . . . . . . . . . . . . . . . . . . 34 Upgrading intermittent water systems and its effect on booster pumps 39 2.4.1 Incremental improvements in intermittent supply . . . . . . . 39 2.4.2 A transition to continuous supply . . . . . . . . . . . . . . . . 40 2.1.3 2.2 . . . . . . . . . . . . . . . . . . . . . . . . . . Effects of intermittent water supplies Prior art 43 3.1 Household contamination prevention devices . . . . . . . . . . . . . . 43 3.2 Self-actuated, full-bore pressure regulating valves . . . . . . . . . . . 46 3.3 Collapsible tubes or ‘Starling Resistors’ . . . . . . . . . . . . . . . . . 48 The iterative design of a self-actuating, full-bore, back-pressure regulating valve 51 9 4.1 4.2 Preliminary concepts and their evaluation . . . . . . . . . . . . . . . 52 4.1.1 Keeping upstream pressure positive . . . . . . . . . . . . . . . 52 4.1.2 Minimizing hydraulic losses 54 4.1.3 Concept generation and selection . . . . . . . . . . . . . . . . 55 Early iterations: searching for stable throttling . . . . . . . . . . . . . 55 4.2.1 Full factorial experiment to determine insert geometry and tube choice 4.3 4.4 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 Alpha prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 4.3.1 Durability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 4.3.2 Withstanding system pressure . . . . . . . . . . . . . . . . . . 69 4.3.3 Tamper resistant 70 4.3.4 Safe for drinking water contact . . . . . . . . . . . . . . . . . 71 4.3.5 Laboratory fatigue test . . . . . . . . . . . . . . . . . . . . . . 71 4.3.6 Field test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 4.3.7 Improvements needed . . . . . . . . . . . . . . . . . . . . . . . 74 Beta prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 4.4.1 Increasing the valve’s toughness . . . . . . . . . . . . . . . . . 77 4.4.2 Preventing set-point drift . . . . . . . . . . . . . . . . . . . . . 79 4.4.3 Detailed design of the beta prototype . . . . . . . . . . . . . . 82 4.4.4 Laboratory fatigue test . . . . . . . . . . . . . . . . . . . . . . 85 4.4.5 Field test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 4.4.6 Improvements needed . . . . . . . . . . . . . . . . . . . . . . . 87 Experimental and numerical methods 89 5.1 89 Experimental methods . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1.1 Laboratory setup and experimental method . . . . . . . . . . 89 5.1.2 Fatigue test setup . . . . . . . . . . . . . . . . . . . . . . . . . 92 5.1.3 Field testing setup and method . . . . . . . . . . . . . . . . . 92 5.2 Hydraulic notation and units . . . . . . . . . . . . . . . . . . . . . . . 98 5.3 Quantifying the intrusion risk 98 . . . . . . . . . . . . . . . . . . . . . . 10 5.4 5.3.1 Water leakage and its pressure dependency . . . . . . . . . . . 100 5.3.2 Quantifying the potential for contaminant intrusion . . . . . . 101 5.3.3 Defining the contaminant intrusion ratio . . . . . . . . . . . . 106 5.3.4 Contaminant concentration at the consumer’s tap . . . . . . . 107 5.3.5 Connecting contaminant intrusion to health risks . . . . . . . 109 Other numerical methods . . . . . . . . . . . . . . . . . . . . . . . . . 110 5.4.1 Monte Carlo sensitivity study 110 5.4.2 Determining upstream system pressure during booster pump events 5.4.3 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 Determining the supply hours of intermittently supplied houses 113 Results 115 6.1 Effect of the valve on a single pump cycle . . . . . . . . . . . . . . . . 117 6.2 Case study 1: applying the metrics to a house in Old Delhi (J02) . . 119 . . . . . . . . . . . . . . . . . . 119 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 6.3 6.2.1 Duration of negative pressure 6.2.2 CIR 6.2.3 Health risk . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 Sources of uncertainty in the data . . . . . . . . . . . . . . . . . . . . 123 6.3.1 Variability in baseline characteristics of customer connections 123 6.3.2 Case study 2: the effects of variability on a house in Lado Sarai (M05) 6.4 6.5 6.6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 Total aggregate impact of the valve . . . . . . . . . . . . . . . . . . . 127 6.4.1 Duration of negative pressure . . . . . . . . . . . . . . . . . . 127 6.4.2 CIR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 Aggregate impact of the valve on continuously-supplied houses . . . . 129 6.5.1 Reduction in the duration of negative pressure . . . . . . . . . 132 6.5.2 Reduction in CIR . . . . . . . . . . . . . . . . . . . . . . . . . 135 6.5.3 Reduction in health risk 138 . . . . . . . . . . . . . . . . . . . . . Aggregate impact of the valve on intermittently supplied houses 6.6.1 Intermittent connections require data filtering 11 . . . 138 . . . . . . . . . 139 6.7 6.8 7 6.6.2 Duration of negative pressure during supply hours . . . . . . . 141 6.6.3 CIR during supply hours . . . . . . . . . . . . . . . . . . . . . 144 6.6.4 Health risk during supply hours . . . . . . . . . . . . . . . . . 148 Valve’s aggregate impact whenever water was being supplied . . . . . 150 6.7.1 Duration of negative pressure . . . . . . . . . . . . . . . . . . 150 6.7.2 CIR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150 6.7.3 Health risk . . . . . . . . . . . . . . . . . . . . . . . . . . . . The necessity of a pilot study and recommendations for its scoping . Conclusions and future work 153 154 157 Bibliography 160 A Measuring the precision and accuracy of the pressure loggers 167 A.1 Measurement method . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 B Other ideas for future work 171 C Impact of the valve on each house 173 12 List of Figures 2-1 Sewage and water pipes cross contaminating . . . . . . . . . . . . . . 29 2-2 Degrading intermittency of supply in four large Indian cities . . . . . 31 2-3 Range of supply hours in select cities in India . . . . . . . . . . . . . 32 2-4 A typical household pipe configuration and why booster pumps are required . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2-5 RBP configuration 1: official booster pump connection . . . . . . . . 35 2-6 RBP configuration 2: booster pump before and after the sump . . . . 36 2-7 RBP configuration 3: booster pump is directly connected without a sump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 3-1 Air gap as a method for stopping negative pressure . . . . . . . . . . 44 3-2 Standard vacuum breaker . . . . . . . . . . . . . . . . . . . . . . . . 45 3-3 Barometric loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 3-4 Self-actuated back-pressure valve with diaphragm . . . . . . . . . . . 47 3-5 Self-actuated full-bore valve . . . . . . . . . . . . . . . . . . . . . . . 48 4-1 Overly complex variable set point valve . . . . . . . . . . . . . . . . . 54 4-2 Bend and constriction losses in some types of commercial valves . . . 54 4-3 Structured brainstorming sketches, part 1 . . . . . . . . . . . . . . . . 56 4-4 Structured brainstorming sketches, part 2 . . . . . . . . . . . . . . . . 57 4-5 Selected concept: a deflating tube as a back pressure valve . . . . . . 58 4-6 Detail of open and closed state of collapsible tube . . . . . . . . . . . 58 4-7 Observed instability in unmodified tube . . . . . . . . . . . . . . . . . 58 4-8 Three prototypes attempting to remove instability . . . . . . . . . . . 60 13 4-9 Optimal insert shape . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 4-10 Optimal insert cross-section geometry . . . . . . . . . . . . . . . . . . 61 4-11 Linear Opening insert geometry . . . . . . . . . . . . . . . . . . . . . 61 4-12 Linear Side Profile insert geometry 61 . . . . . . . . . . . . . . . . . . . 4-13 Total seconds of flutter for each design parameter . . . . . . . . . . . 62 4-14 Minimum upstream pressure allowed for each design parameter . . . . 62 4-15 Bill of materials and exploded view of alpha prototype . . . . . . . . 65 . . . . . . . . . . . . . . . . . . . . . . . 66 4-17 Alpha prototype insert part drawing; units of inches . . . . . . . . . . 67 4-18 Alpha prototype’s strain transition layers . . . . . . . . . . . . . . . . 68 4-19 Detail of alpha prototype’s two possible external support mechanisms 70 4-16 3D CAD of alpha prototype 4-20 Process steps in the outsourced casting of the urethane inserts for the alpha prototype . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 4-21 Identifying where fatigued prototypes leaked . . . . . . . . . . . . . . 73 4-22 Detail of failure location and suggested cause in fatigue test . . . . . 73 . . . . . . . . . . . . . . . . . . . 74 4-23 Alpha prototypes after field testing 4-24 Alpha prototypes’ field test required a flexible hose to reduce the shear load on the valve; from bottom to top, pressure logger, alpha prototype, check valve to air, flexible hose, booster pump . . . . . . . . . . . . . 4-25 55% of alpha field units failed in shear at the threads . . . . . . . . . 75 76 4-26 Alpha prototype’s set-point drift because of a slow leak into the reference chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 4-27 The peel force was dramatically reduced by turning the membrane inside-out at each end of the middle layer . . . . . . . . . . . . . . . . 80 4-28 Section views of the beta prototype . . . . . . . . . . . . . . . . . . . 81 4-29 Beta prototype’s exploded view . . . . . . . . . . . . . . . . . . . . . 82 . . . . . . . . . . . . . . . . . . . . 83 . . . . . . . . . . . . . . . . . . . . . . . 83 4-32 Beta prototype’s middle layer part drawing . . . . . . . . . . . . . . . 84 4-33 Field modification of the beta prototype 87 4-30 Beta prototype’s assembly stages 4-31 Beta prototype’s outer pipe 14 . . . . . . . . . . . . . . . . 5-1 Laboratory test setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 5-2 Laboratory experimental setup photo . . . . . . . . . . . . . . . . . . 91 5-3 Layout of the fatigue test; ‘R’ are the two regulators; pressures were typically set to 10m and 17m 5-4 . . . . . . . . . . . . . . . . . . . . . . 93 Built fatigue testing apparatus; blue tubes carry 17m pressure for the inside of the valves; pink tubes carry 10m pressure for the outside of the valve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 5-5 Pressure loggers from Hydreka . . . . . . . . . . . . . . . . . . . . . . 94 5-6 Field test schematic and predicted pressure performance . . . . . . . 95 5-7 Three prototypes attempting to remove instability . . . . . . . . . . . 95 5-8 Field equipment arrangement for beta prototype; currently taking control measurements — valve is bypassed . . . . . . . . . . . . . . . . . 96 Spacial modeling of contaminants near leaky pipes . . . . . . . . . . . 100 5-10 Pressure drop along the service connection . . . . . . . . . . . . . . . 104 5-11 Two examples of upstream pressure estimation . . . . . . . . . . . . . 112 5-12 Determining supply times . . . . . . . . . . . . . . . . . . . . . . . . 114 6-1 Effect of the valve on different types of house connections . . . . . . . 118 6-2 Negative pressure reduction example from J02 in Old Delhi . . . . . . 120 6-3 Contamination risk of J02 . . . . . . . . . . . . . . . . . . . . . . . . 122 6-4 How the valve caused a longer duration of negative pressure at M05 . 125 6-5 Evidence of a mean shift in the sensor at M05 . . . . . . . . . . . . . 126 6-6 Severe negative pressure reduction . . . . . . . . . . . . . . . . . . . . 128 6-7 Inconclusive impact of duration of negative pressure . . . . . . . . . . 130 6-8 Inconclusive impact on CIR . . . . . . . . . . . . . . . . . . . . . . . 131 6-9 Time spent at negative pressure on continuous connections . . . . . . 132 5-9 6-10 Detail of pressure history at J07; highlights the anomaly during the control measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 6-11 Time spent at negative pressure on continuous connections; adjusted for J07 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 134 6-12 Accounting for the pressure anomaly in M01 . . . . . . . . . . . . . . 135 6-13 Evidence of a drift in the valve’s set point at M01 . . . . . . . . . . . 136 6-14 Valve’s ability to change CIR values for continuously pressurized houses 137 6-15 Variability in the supply pressure to intermittent houses was larger than the effect of the valve . . . . . . . . . . . . . . . . . . . . . . . . 141 6-16 Inconclusive impact on negative pressure duration on intermittentlysupplied houses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-17 Inconclusive impact on total CIR on intermittently-supplied houses . 142 143 6-18 Negative pressure reduction during supply hours on intermittent houses 145 6-19 Unusual performance of M04 . . . . . . . . . . . . . . . . . . . . . . . 146 6-20 CIR reduction during supply hours on intermittent houses; red dots are expected values, red error bars are 99% range and blue error bars are the 95% range . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-21 Reduction of negative pressure during supply times 6-22 Reduction of CIR during supply times 147 . . . . . . . . . . 151 . . . . . . . . . . . . . . . . . 152 A-1 Accuracy and precision or pressure loggers . . . . . . . . . . . . . . . 168 C-1 Pressure summary: Azad Market: J01 . . . . . . . . . . . . . . . . . 174 C-2 Pressure summary: Azad Market: J02 . . . . . . . . . . . . . . . . . 175 C-3 Pressure summary: Azad Market: J03 . . . . . . . . . . . . . . . . . 176 C-4 Pressure summary: Azad Market: J04 . . . . . . . . . . . . . . . . . 177 C-5 Pressure summary: Azad Market: J05 . . . . . . . . . . . . . . . . . 178 C-6 Pressure summary: Pitampura: J06 . . . . . . . . . . . . . . . . . . . 179 C-7 Pressure summary: Pitampura: J07 . . . . . . . . . . . . . . . . . . . 180 C-8 Pressure summary: Vivek Vihar: J08 . . . . . . . . . . . . . . . . . . 181 C-9 Pressure summary: Pitampura: J09 . . . . . . . . . . . . . . . . . . . 182 C-10 Pressure summary: Pitampura: J10 . . . . . . . . . . . . . . . . . . . 183 C-11 Pressure summary: Pitampura: J11 . . . . . . . . . . . . . . . . . . . 184 C-12 Pressure summary: Pitampura: J12 . . . . . . . . . . . . . . . . . . . 185 C-13 Pressure summary: Vasant Vihar: M01 . . . . . . . . . . . . . . . . . 186 16 C-14 Pressure summary: Vasant Vihar: M02 . . . . . . . . . . . . . . . . . 187 C-15 Pressure summary: Shivalik: M03 . . . . . . . . . . . . . . . . . . . . 188 C-16 Pressure summary: Shivalik: M04 . . . . . . . . . . . . . . . . . . . . 189 C-17 Pressure summary: Lado Sarai: M05 . . . . . . . . . . . . . . . . . . 190 C-18 Pressure summary: Lado Sarai: M06 . . . . . . . . . . . . . . . . . . 191 C-19 Pressure summary: Lado Sarai: M07 . . . . . . . . . . . . . . . . . . 192 17 THIS PAGE INTENTIONALLY LEFT BLANK 18 List of Tables 2.1 Water balance definitions . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.2 CPHEEO static pressure requirements . . . . . . . . . . . . . . . . . 38 4.1 Initial functional requirements and design parameters . . . . . . . . . 52 4.2 Logic of a back-pressure regulating valve with a set point of X . . . . 52 4.3 Design variables considered in search of stable operation . . . . . . . 59 4.4 Insert geometry test matrix; repeated for silicone and latex soft tube 61 4.5 Alpha prototype’s functional requirements and design parameters . . 64 4.6 Alpha fatigue test matrix: tube material and inner fillet radius . . . . 72 4.7 Beta prototype’s functional requirements and design parameters . . . 78 4.8 Beta fatigue test matrix: clearance for membrane and the insert’s inner fillet radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 5.1 Laboratory measurement equipment . . . . . . . . . . . . . . . . . . . 91 5.2 House connection characteristics . . . . . . . . . . . . . . . . . . . . . 97 5.3 The assumptions behind Equation 5.14 . . . . . . . . . . . . . . . . . 108 5.4 The effects of Equation 5.14’s assumptions . . . . . . . . . . . . . . . 108 5.5 WHO’s example thresholds for E. coli contamination risk . . . . . . . 110 5.6 Assumptions required to connect CIR values to health risks . . . . . . 110 5.7 Allowable IC and CIR levels . . . . . . . . . . . . . . . . . . . . . . . 111 5.8 Measured supply times . . . . . . . . . . . . . . . . . . . . . . . . . . 113 6.1 Summary of negative pressure reduction on continuously supplied houses136 6.2 Summary of contamination reduction on continuously supplied houses 19 137 6.3 Reduction in health risk at continuously supplied houses . . . . . . . 138 6.4 Booster pump induced pressure characteristics on intermittent houses 140 6.5 Duration of negative pressure at intermittently-supplied houses during supply hours . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.6 Summary of contamination reduction on intermittently supplied houses, during supply hours . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.7 148 148 Reduction in health risk at intermittently supplied houses, during supply hours . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 6.8 Summary of negative pressure reduction during supply hours . . . . . 151 6.9 Summary of contamination reduction during supply hours . . . . . . 153 6.10 Reduction in health risk during supply hours . . . . . . . . . . . . . . 153 6.11 Required supply pressure to limit the valve’s flow reduction . . . . . . 154 A.1 169 Total variation in the logging equipment 20 . . . . . . . . . . . . . . . . Nomenclature ππΌ 3 Average Daily Intrusion Volume [π /πππ¦ ] ππΏ 3 Daily averaged leakage rate for a given section of pipe [π ] π The distance along the supply pipe at which pressure becomes negative [π] π΄eq 2 Equivalent area of a single orifice to account for all leaks [π ] πΆπ»π Hazen-Williams constant π· Pipe diameter [m] π Friction factor [unitless] π 2 Gravitational acceleration [π/π ] π»loss Frictional pressure head loss [m] π»π The contaminant pressure, positive if it forces them inside the pipe [m] π»π Pressure in the pipe [m] π»π Supply pressure at the distribution main [π] π»π΅π Pressure at the suction inlet of the booster pump [π] π»π π Duration of water supply [hours/day] πΏπΆ 2.5 Empirical leakage coefficient [π /π ] πΏπ The length of pipe [m] 21 πΏπΏ Leakage level [%] π 3 Volumetric flow rate [π /π ] ππΌ 3 The volumetric flow rate of contaminant intrusion into the pipe [π /π ] ππΏ Volumetric flow rate of leaking water π Duration of Observation [s] π‘ Elapsed Time [s] π′ The duration of either the control or treatment measurement at a given house [days] ππ The total duration of the control and treatment trials at a given house [days] ππΌ′ (π‘) The relative total volume of intruded contaminants over a given time interval [π 3 /πΏπΆ ] ππΏ′ (π‘) The relative total volume of leakage over a given time interval [π π₯ Length along pipe from supply towards consumer’s house [m] CFU A measure of contamination; colony forming units [/100mL] CIR Contaminant Intrusion Ratio [unitless] IC Intrusion concentration V Average fluid velocity [m/s] 22 3 /πΏπΆ ] Chapter 1 Introduction This thesis began with a desire to make a contribution to improving the quality of water delivered to the 279 million people in India who have access to piped water on their premises [1]. Most water distribution systems in India operate intermittently [2, 3], which makes them more likely to recontaminate the distributed water than continuously-operated water systems [4, 5]. As part of the MIT Tata Center for Technology and Design, the author traveled to India on five separate occasions and met with water-system designers, operators, and users in Mumbai, Bangalore, and New Delhi. A pivotal moment in the project’s scoping occurred when one system engineer admitted that he only knew what the average pressure in his system was when the electricity was not working. While the city’s electricity was working, households would connect small pumps directly to the water network, rapidly withdrawing water and reducing system pressure. However, when the electricity was not working, these pumps could not be used and the system pressure increased to a known level. To increase withdrawal rates, these residential booster pumps create a negative pressure in the pipe leading to the house; unfortunately, this negative pressure increases the risk of contamination. Rapid withdrawal rates lead to dramatic reductions in water pressure at the extreme ends of the water system and lead to unequal distribution of water [6]. To protect households from contamination and to protect the water system from low and unequal pressure, it is typically mandatory in India to install an isolation tank, 23 called a sump, between the city’s water supply pipe and any residential booster pump. A more detailed consideration of the connection between contamination, pressure, and booster pumps can be found in Chapter 2. Some cities are better at enforcing the prohibition against using a booster pump without a sump than others. Specifically, because of population density and political precedent, the regulations regarding sumps have not been enforced in New Delhi. Therefore, an alternative way of protecting customers from contamination, and protecting New Delhi’s water system from low and unequal pressure is required. Existing options for such protection were reviewed for their applicability to Delhi’s water system and were found to be either easily tampered with or too costly; a brief summary of such prior art is included in Chapter 3. Working in collaboration with New Delhi’s water utility, the Delhi Jal Board (DJB), a patent-pending, self-actuating, full-bore, back-pressure regulating valve was developed and tested. The valve’s simple mechanism relied on a collapsing tube that would close when negative pressure was present inside the pipe. The major innovation in this valve’s design was a specially-shaped insert which prevented oscillations, which are typically found in such collapsing tubes. This stabilized-collapsing-tube mechanism was encapsulated in a rigid valve shell. The valve’s design and the design’s evolution are documented in Chapter 4. The experimental and numerical methods for testing and interpreting the valve’s performance, presented in Chapter 5, were used to measure the valve’s performance in laboratory flow tests, fatigue tests, and field demonstrations. Field tests of two design iterations were conducted in New Delhi in January and March of 2014; the successful results of these tests are presented and discussed in Chapter 6. While the original application for the valve designed herein was New Delhi’s water utility, it holds potential for any water system in which customers use booster pumps without isolation tanks, or any local where it might be desirable to replace isolation tanks with an equivalent valve that is smaller and cheaper. Further conclusions and opportunities for future work are presented in Chapter 7. 24 Chapter 2 Background While the bulk of this thesis focuses on the effects of residential booster pumps and a water valve to reduce the risk of contamination, this section explores the issue of water contamination from first principles. First, it considers the mechanism of contamination in water distribution systems; second, it specifically explores the consequences of intermittent water supply; third, it addresses how booster pumps are typically installed and the effects of these installation method; and finally, this chapter concludes with a description of the two current and competing opinions on how to best address the problems associated with intermittent water distribution, noting that in until water is distributed continuously and at uniformly high pressure, the problem of residential booster pumps must be addressed. 2.1 Contaminant intrusion The intrusion of contaminants into water distribution systems has caused significant public health concerns and several historical crises [7]. Intrusion requires three things [8]: 1. a pathway for contaminants to enter the water system, 2. contaminants to be present in the vicinity of the pathway, and 3. a pressure gradient such that the contaminants flow into the water system. 25 Table 2.1: Water balance definitions, unaccounted for water comprises of real losses and apparent losses; adapted from [9] This section will sequentially evaluate whether each of these requirements is met in intermittent water systems in general, in India, and then specifically in New Delhi. 2.1.1 Leakage rates indicate the pathway size for contaminant intrusion The first of three requirements for contaminant intrusion to occur is that a pathway for contamination must exist. Were a physical pathway to exist between the inside of a water pipe and the outside, one would expect that most of the time, such a pathway would have clean water leaking out of it. The proxy of leakage rates is therefore used to assess the size and location of potential pathways for contamination. While all water systems leak, leakage levels in water systems in India are typically extremely high and therefore many pathways for contamination are present. Because the exact level of physical leakage in a water system is difficult to measure, many proxy indicators are used; their overlapping definitions are captured in Table 2.1. Unaccounted for water (UFW) is a frequently used metric because it is easily calculated, but it obscures the magnitude of physical leakage by combining it with inaccurate water meters and unauthorized connections. Another common metric is non-revenue water (NRW), which further obscures physical leakage by including metered but unpaying customers. The WHO and UNICEF have estimated an average level of UFW in urban water systems across Asia, Africa, and South and Central America as 40% [10]; within this UFW, they estimate that physical losses comprise the majority. 26 In India in 2007, self-reported UFW levels of 20 water utilities averaged 32% [3]. In New Delhi in 2011, non-revenue water was reported by Delhi’s water utility, the Dehli Jal Board (DJB), as 52% [11]. Delhi’s physical losses were estimated to be 20-25% [11]. For comparison, New York and Toronto average UFW levels of 15% [12]. A simple percentage comparison however, hides how large the problem is. Cities in India on average supply water for only four hours per day, often at low pressure [3]. To equitably compare leakage rates across different cities, the leakage rates must be scaled according how long each system was pressurized and at what pressure. Using a simple leak dependency where flow is proportional to the square root of pressure [13], we find Equation 2.1 (for a more thorough treatment of leakage equations see Section 5.3.1, for more information about pressure notation and units see Section 5.2). Where πΏπΏ is the leakage level in percent, and π» π»π π is the number of hours per day of supply, is the average pressure during supply times. πΏπΏadjusted (π»π π)adjusted = πΏπΏold (π»π π)old √οΈ π»adjusted π»old (2.1) Converting the average leakage levels in India to their equivalent 24-hour amounts would make them six times higher. Taking New Delhi’s leakage rate as an example, if one optimistically assumes an average supply of 6 hours per day at 7m of pressure, then to keep those sames pipes pressurized to 17m for 24 hours per day, would create a leakage level of 125%, i.e. it would require 125% of the current total water usage to go towards leakage alone. Adding customer demand to this calculation, it would therefore require twice the current amount of water to supply all customer needs and to keep the New Delhi’s current pipe system continuously pressurized. It must therefore be concluded that in existing water systems in India, significant pathways for contaminant intrusion exist. While upgrading these systems would reduce many of the pathways for contamination, some leakage will always exist and so the pathway for contaminants can never be completely removed. Keeping water safe from contaminant intrusion must therefore rely on avoiding one or both of the 27 other two requirements for contaminant intrusion. 2.1.2 Contaminants may be present because of the unplanned locations of water pipes Eliminating the presence of contaminants is likely infeasible. Officially, Indian standards insist on three meters of horizontal separation between a water pipe and sewers or other contamination sources. When water pipes need to crossover a contamination source, the water pipe must be at least 0.5m above the contamination source [14, p.388]. Even where safety distances are maintained, during and after a rainfall, soils can become saturated with rainwater. In one detailed field study of water contamination in an Indian city, water contamination was found to increase significantly immediately after major rainfalls [4]. To make matters worse, in most cities in India, the drainage, sewer, and water pipe networks have been developed independently, haphazardly, and very shallowly, preventing adherence to the required separation. One field engineer even reported that plumbers will often route the water pipe through (not over or under) an open channel sewer to make the job go faster. A typical example of this cross-contamination potential is shown in Figure 2-1. It therefore seems inevitable that contaminants are present in the vicinity of water pipes; if intrusion is to be avoided, another prevention strategy must be taken. 2.1.3 Pipe pressure can deter contamination Where a pathway exists and contaminants are present, pipe pressure is the last line of defense against contaminant intrusion. Intrusion can only occur when the con- taminants outside the pipe have a higher pressure than the water inside the pipe. This occurs either when the pressure in the pipe goes below atmospheric (negative pressure), or when the contaminants around the pipe become positively pressurized. Low water system pressure led to water contamination in eight case studies compiled by Lee and Schwab [7]. Contaminant pressure is usually less than the depth of 28 Figure 2-1: A leaking sewage pipe drips onto an array of household water service connection pipes, ensuring contaminants are in present around the pipe the pipe underground and can occur when the pipe has been laid below the ground water level (unlikely in urban centers in India), when soils become temporarily saturated with rainwater, or when sewers plug and the sewage becomes pressurized. One Indian engineer reported that the typical first response to customer contamination complaints is to check to see whether a sewer near the customer’s house has become plugged. To overcome the risk of contamination, distribution systems should maintain a minimum safety pressure greater than the expected maximum external contaminant pressure. Worse than low pressure, is no pressure. While pipes are unpressurized, they have no defense against the intrusion of gravity-driven, percolating, contaminants. Intermittent water systems, are by definition, frequently unpressurized. On average, water systems in India are unpressurized for 20 hours per day [3]. When these systems are turned on, the initial water flowing through them must flush out the contaminants which have been filling the system for the past 10 to 20 hours. These contaminants typically exit the system through consumers’ taps. Kumpel and Nelson found that 29 contamination indicator organisms were present in concentrations that were an order of magnitude higher during the initial flushing period than during steady-state operations [4]. Compounding this quality concern, many consumers use an overhead storage tank which fills during supply hours. Such houses are often plumbed to directly fill the tank whenever there is water pressure, mixing therefore the lower quality flushing water with the higher quality steady-state water. While not within the scope of this thesis, creating a new system for automatically storing the initial lower quality water separately from the cleaner water represents a possible improvement to the status quo. The most serious source of contamination however, comes from negatively-pressurized pipes. Negative gauge pressure by definition means that the pressure inside the pipe is less than atmospheric pressure; it therefore acts to actively pull contaminants into the pipe. Section 5.3 develops the required theory to quantify the relative risks of unpressurized and negatively-pressurized systems. Given the difficulty in preventing the first two requirements for contaminant intrusion, positive pressure must be maintained. The impossibility of maintaining positive pressure in intermittently supplied water is one of the main reasons many in this sector advocate for a transition back to continuously operated (24/7) water distribution. 2.2 Intermittent water systems Despite the contamination danger, in Africa, Latin America, and the Caribbean, onethird of distribution systems operate intermittently [12]; in Asia, half do [12]; and, most major water systems in India do [3]. These distribution systems were neither intended to be nor designed to be operated intermittently [6], but are forced to do so because the combination of leakage rates and consumer demand exceeds the available water supply. Water suppliers forced to adopt the strategy of intermittent water supply, might initially only need to shut the system down for a few hours per day. Consumers would withdraw a similar volume of water as during continuous supply (only at a 30 Figure 2-2: Historically, maintaining an intermittent system that provides water for the majority of the day has not been sustainable, reproduced from [15] faster rate), but the leakage levels would be reduced, according to Equation 2.1. This reduction in leakage level would make up the supply deficit. The longterm viability of this strategy however, is dubious; historically the amount of time a utility in India has been able to supply water has degraded over time, as set out in Figure 2-2, reproduced from [15]. While the mechanism for this degradation is debatable, the historical viability of turning off the system for only a few hours per day in India is not. Therefore, most cities in India supply water for approximately four hours per day or less, as summarized in Figure 2-3. 2.2.1 Effects of intermittent water supplies In addition to the primary concern of higher contamination risks, when compared to continuous water supply, intermittent water supply does the following: 1. distributes water less equitably, 2. causes the water network to degrade more quickly, and 3. passes on more adaptation costs to the consumer. High frictional losses during supply times create severe geographic inequalities in the pressure and duration of water supply [6, 16], which disadvantage consumers at 31 Figure 2-3: Most cities in India supply water for approximately four hours per day or less, reproduced from [3, p.21] the extremities of the water system whose pressure would already have been lower than customers closer to supply reservoirs. This phenomenon arises because the total consumption of water after a transition to intermittent water supply does not change dramatically. Demand is therefore concentrated in shorter periods of time, and so flow rates increase. As pressure loss is approximately proportional to the flow rate squared [13], even a reduction to 12 hours of supply per day which could lead to a doubling of flow rates, would translate into a quadrupling of pressure losses. In the filling and draining of water systems, the sudden change from open-channel flow to pressurized flow can create large pressure transients and cause accelerated pipe wear and bursts [17, 18]. While in continuously-operated systems these dangerous events occur infrequently, they can happen as often as twice per day in intermittently operated systems. One water system engineer in India reasoned that since most consumer taps are left open in intermittent systems, they act as distributed air release valves that dampen this effect. Transient expert Bryan Karney however, was skeptical 32 that this would provide significant damping [19]. This assertion will be verified in future work as a possible opportunity for improving the quality and durability of the existing water networks. Nevertheless, it remains true that the intermittent operation of a water system accelerates its degradation. Previous studies have focused in detail on the adaptations that consumers are forced to make in response to lower pressure and shorter supply times as a method of quantifying the indirect costs of intermittent supply to consumers [20, 21, 22, 23]. These studies often aimed at determining how much households would be willing to pay for the provision of higher quality of water. Such lists of consumer adaptations often include: 1. water filtration or boiling to address the lower quality; 2. time spent collecting water, or waiting for the system to turn on; 3. purchasing alternative (and possibly lower quality) water to supplement supply; 4. a storage tank (often on the roof ) to make water available during non-supply hours; and sometimes 5. a residential booster pump to supplement the system pressure and/or increase the flow rate. This final adaptation where consumers use residential booster pumps is of special relevance to this thesis and is the focus of the next section. 2.3 Residential booster pumps A residential booster pump (RBP) is a centrifugal pump in the home of a water consumer. They are typically used when consumers wish to supplement either the pressure or the flow rate provided by the water utility. They range in size from 0.5 to 1.5 horsepower (hp), with the most common being 0.5 and 1hp. Some methods of connecting the RBP can lead to an increased risk of contamination. This section first 33 Storage Tank No flow Supply Pressure Pressure Low flow Moderate flow Water Meter M Ground Level Distribution Main 2 Story Home Service Connection Pipe Figure 2-4: A household connection without a RBP overlaid with the pressure profile for house connections at different stories; upper stories and the roof tank do not receive acceptable flow rates and require a booster pump details how pumps are typically configured and then elaborates on the risks associated with these configurations. 2.3.1 Booster pump configurations Figure 2-4 depicts how a house would typically be connected to a high pressure water supply without a RBP. The pipe’s pressure profile is overlaid on the diagram, where the vertical axis is the pressure head and the horizontal axis is the location along the pipe. The slope of this pressure profile indicates the direction and magnitude of flow; water flows in the direction of decreasing pressure in proportion to the square root of the slope. When the supply pressure is less than the height of the building, residents of upper stories receive little or no water. To overcome this problem, consumers typically use RBPs. RBPs can be connected in one of three ways: 1. Figure 2-5 shows the officially-sanctioned configuration. After passing through the water meter, the water discharges into an underground reservoir called a sump. The booster pump draws water from the sump and pushes it into the overhead tank. Note that the booster pump can only draw water from the sump and therefore the RBP neither affects the flow rate into the sump, nor creates a negative pressure in the service connection. 34 Tank Pressure Supply Pressure Moderate flow Water Meter Figure 2-5: RBP configuration 1: P M Ground Level Distribution Main 2 Story Home Service Connection Pipe Sump the officially sanctioned method of connecting a booster pump; the pump only draws water from the sump and cannot create a negative pressure in the service connection 2. Figure 2-6 shows a modification of the “official” configuration. Where sumps exist but consumers are not content with the rate at which the sump fills, a second booster pump is often installed upstream of the sump. The sole purpose of this upstream pump is to increase the flow rate into the sump. Since the upstream booster pump is connected directly to the service connection pipe, it increases flow by creating a negative pressure at its inlet. This negative pressure spreads upstream in the service connection, represented by the shaded orange region in Figure 2-6. 3. Figure 2-7 shows another common configuration. Where houses do not have sumps, the pump connects the service connection pipe directly to the overhead tank. This configuration can also create negative pressure upstream of the booster pump. Configurations 2 and 3 are problematic because each can induce negative pressure, which creates four connected problems: 1. it increases significantly the risk of contamination as discussed in Section 2.1.3; 2. it increases flow rate which, when many people have booster pumps, further lowers the system pressure and increases distribution inequality; 35 P Supply Pressure Pressure M Tank = Booster Pump = Water Meter 2 Story Home Very large flow M Ground Level Distribution Main Service Connection Pipe P P P<0 Sump Figure 2-6: RBP configuration 2: a booster pump is located before the sump and fills it faster; the pump after the sump fills the overhead tank; the flow rate is increased by creating a negative pressure which also increases contamination risk P Supply Pressure Pressure M Tank = Booster Pump = Water Meter 2 Story Home Very large flow M Ground Level Distribution Main Service Connection Pipe P P<0 Figure 2-7: RBP configuration 3: booster pump is directly connected without a sump; this configuration can also cause negative pressure 36 3. it creates pump wars, in which neighbors with bigger pumps receive more water; and 4. it prevents system operators from being able to use water pressure as a means of controlling customer flow rates. Wherever intermittent systems exist, booster pumps are likely to be found [21, 22, 24, 25]. Specifically in India, their use is frequent [20, 21, 22]; however local governments vary in their insistence on the use of the “official” configuration style. This thesis’s geographic focus on New Delhi arose because of Delhi’s unique position regarding booster pumps. During approximately 40 house inspections in New Delhi by the author, configurations 2 and 3 were frequently observed, with a ratio of approximately 1:2. Configuration 1 was observed in New Delhi only once. Despite their widespread use, because of the severe consequences of configurations 2 and 3, the Delhi Jal Board Act has specifically prohibited the connection of residential booster pumps to the service connection pipe [26, §19(3)]. The DJB is however mandated by the Indian Central Public Health and Environmental Engineering Organization (CPHEEO) to supply pressure at the point where the service connection meets the distribution main (the location of a check valve called a ferrule ) in accordance with the Table 2.2. Because of municipal building codes in Delhi and the high population density, water must be supplied at 17m of pressure in order to ensure that apartments on the third story (Indian second story) can receive water. Since the DJB cannot meet this requirement, enforcing the prohibition against booster pumps has been described as politically infeasible; enforcing the prohibition would cut off consumers from water. Could these connections at least be forced to construct sumps? In densely populated areas, retroactively constructing sumps would be prohibitively expensive. A section on booster pumps would not be complete without three quotes from upper managers at the Delhi Jal Board which summarize the practitioners’ perspective on the problem: “It [the problem of booster pumps] f_cks my hydraulics” 37 Table 2.2: CPHEEO static pressure requirements, from [14, p.361 §10.3.3] Height of Building Required Static Pressure Single story 7m Two story 12m Three story 17m “Booster pumps are like a disease [once one person has one, they spread]” “Booster pumps are our biggest problem” New Delhi’s water system, therefore, required a method for making booster pump configurations 2 and 3 safer and more equitable without the space or cost of an underground sump. More specifically, this could only be accomplished if booster pumps were prevented from inducing negative pressure in the service connection pipe upstream of them. Therefore, the key objective of the design presented in this thesis was to limit the ability of booster pumps to create negative pressure in the service connection pipe, thereby decreasing contamination and increasing distribution equality. Should a solution to this problem be found, it would have applications beyond just New Delhi; were it to become an approved alternative to configuration 1, households across India would have a smaller and cheaper way of safely connecting their booster pumps. A solution to this problem would, therefore, benefit households and the water utility in New Delhi, in other intermittently-supplied cities in India, and in any intermittently-supplied city that uses booster pumps. During the development of this project, however, a number of practitioners questioned the long-term relevance of the problem of booster pumps given the current policy push for intermittent water systems to convert to continuous water systems. The final section in this chapter seeks to review the current strategies for improving intermittency and to determine the likelihood that each strategy would mitigate the problem of booster pumps. 38 2.4 Upgrading intermittent water systems and its effect on booster pumps To address the concerns of water quality and distribution equity associated with intermittently supplied water, two kinds of solutions have been proposed: the first proposes to create incremental improvements in intermittent systems, and the second proposes methods to eliminate intermittency altogether. The author’s conversations with many practitioners in the Indian water sector have suggested that the latter approach has captured much of the policy focus. Where incremental improvements are being made to intermittent water systems and where the transition to high pressure and continuous supply has not yet been completed, the problem of booster pumps is likely to persist; in such circumstances, solutions to the booster pump problem will continue to be relevant. 2.4.1 Incremental improvements in intermittent supply Advocating for the incremental option, a leading expert in intermittent water systems, Professor Vairavamoorthy, has argued that intermittent systems are inevitable and therefore finding better ways of operating them must be a priority [27, p.191]. To mitigate equity concerns he recommended: “Equitable distribution of the limited quantity of water is the keystone of the entire design process...and is a non-negotiable design objective. This objective is achieved mainly through effective pressure management.” To mitigate quality concerns, Vairavamoorthy has advocated for more detailed modeling of contaminants and intrusion [28, 29]. This would allow for prioritized incremental improvements in water safety by, for example, replacing high-risk pipes or lining open sewers that cross many water connections. Residential booster pumps installed in configurations 2 and 3 threaten Vairavamoorthy’s recommended improvements to quality and equity because during supply times, booster pumps actively increase the contamination risk and increase flow rates, 39 which ultimately makes effective pressure management more difficult. The incremental improvement approach, therefore, would not negate the need to safely connect booster pumps; instead it stands to benefit from safely-connected booster pumps. 2.4.2 A transition to continuous supply From the currently dominant perspective, many, including the World Bank’s Water and Sanitation Program (WSP), have advocated for a complete transition back to continuously supplied systems [30], thereby addressing the quality and equity concerns of intermittent water supply. The WSP has demonstrated this approach in three projects in pockets of cities in Karnataka. These pockets successfully made the transition to continuous water supply and found that water quality went up [4], while the total water requirement went down [30]. During the execution of these three demonstration projects in Karnataka, it was found to be more economical to replace the whole water systems than to follow through on the originally planned replacement of 60% of the networks. While the results from this demonstration look positive, it remains unclear as to how feasible such a project would be without external support and oversight from the WSP. Building on the success of these demonstration projects, several cities in India, including New Delhi and Nagpur, have undertaken major public-private partnerships (PPPs) to attempt to transition sections of their distribution networks from intermittently- to continuously-operated. Most of these PPPs do not plan to create an entirely new distribution network, but rather to replace significant portions of the old one and thereby incrementally transition to 24/7 water provision. An incremental transition to continuously-supplied water would require extensive leak repair efforts. Taking New Delhi as an example, if only 7m of pressure and continuous supply were desired, to maintain the current leakage rate of 20%, using Equation 2.1, a 75% reduction in the amount of leaks in the current system would be required. Making the transition to 24/7 and 17m of pressure, would require an 84% reduction in the amount of leaks in the system. To address 84% of leaks, which have unknown locations within the system, would likely be more difficult than a total 40 system replacement, which explains why this was done in the WSP project. When a water system supplies high-pressure water continuously, the problem of booster pumps will be negated; residentially-sized booster pumps cannot create flow rates high enough to create a negative pressure while the supply pressure is 17m; therefore, they will be rendered safe, independent of their connection configuration. However, before and during the transition to continuous high-pressure supply, booster pumps will continue to threaten water quality and distribution equity. Designing and testing a device to limit these consequences that is smaller and cheaper than an underground sump is the focus of this thesis. 41 THIS PAGE INTENTIONALLY LEFT BLANK 42 Chapter 3 Prior art This thesis sought to improve the water quality and distribution equity of urban water distribution networks in India by eliminating the ability of booster pumps to create negative pressure in the service connection pipe. To achieve this aim, the design process detailed in Chapter 4 identified the specific need for a simple and inexpensive, self-actuating, full-bore, back-pressure regulating valve. To contextualize the designs of Chapter 4, this chapter summarizes the existing household contamination prevention devices; it then reviews existing self-actuated, full-bore, and back-pressure regulators; and, finally it concludes with a brief discussion of a collapsing tube, which was the critical component of the final valve design. 3.1 Household contamination prevention devices In continuously-supplied water systems, one possible source of contamination comes from cross-connections. A cross-connection is any connection in the water distribution system, typically in a customer’s home, where the supply of potable water comes in contact with a non-potable liquid, as when, for example, a garden hose dilutes a bucket of liquid fertilizer. If the potable water pressure were to temporarily become negative, contaminants could be drawn back (back-flow) into the potable water supply. To address this public health risk, the US EPA has published the Cross-Connection Control Manual [31]. While this contamination mechanism is different than the one 43 Figure 3-1: An air gap offers excellent protection from negative pressure, but introduces a potential source of contamination and pressure loss; reproduced from [32] caused by booster pumps, both are caused by negative pressure, and therefore, some of the safety equipment designed to prevent contamination from cross-connections was considered as possibly applicable to the problem of booster pumps. Specifically, three devices recommended by the US EPA were reviewed as part of this section: an air gap, a vacuum breaker, and a barometric loop. The US EPA commended the use of an air gap, as in Figure 3-1, as “extremely effective” for the prevention of back-flow [32]. This is the same mechanism employed in the mandatory isolation tank in Indian water systems, referred to there as a ‘sump.’ This solution, however, poses three possible problems: first, the space for such a tank may not always be present; second, the tank creates the possibility for contamination to enter the water supply; and finally, all of the water pressure entering the tank is lost. In India, where an air gap is used, it is typically in a large underground reservoir that can be difficult to clean and costly to build. Another safety mechanism recommended by the US EPA was a vacuum breaker, one such device is shown in Figure 3-2. When vacuum (negative) pressure is present upstream from the device, the vent is pulled open, air enters the system, and the vacuum is broken, i.e. negative pressure was removed. While this device is very effective at preventing back-flow, the problem of booster pumps has different functional requirements. Customers with booster pumps may occasionally wish to create a negative pressure so as to increase their flow rate; the safety device must prevent this and therefore must be difficult to tamper with. The vacuum breaker in Figure 3-2 44 Figure 3-2: A vacuum breaker can prevent negative pressure, but it is easy to tamper with and introduces air into the system; reproduced from [32] could be easily defeated by blocking its vents. Another contextual difference between back-flow and booster pumps is that some houses use booster pumps that are not only above-ground level, but above the water pressure level. For example, an apartment might use a booster pump located 12m above the ground to draw water from the supply pipe which had 10m of supply pressure. To operate, the booster pump would create a negative pressure for several meters below it in the above-ground supply pipe. This negative pressure would not cause contaminant intrusion since it would only occur above the ground where the presence of contaminants is extremely unlikely. Therefore the ideal safety device would only prevent negative pressure in the supply pipe below ground, and would allow elevated pumps to continue to operate. The US EPA’s style of vacuum breaker would introduce air into the system, which would cause an elevated booster pump to lose its prime and the envisioned apartment would receive no water; this option was therefore not pursued. The final potentially relevant device recommendation by the US EPA was a barometric loop, shown in Figure 3-3. If water with a negative pressure enters the bottom of the barometric loop, by traveling upwards its pressure is further reduced and it will eventually vaporize. Once vaporized, the two sides of the loop are no longer hydraulically coupled and the negative pressure on one side of the loop cannot exert 45 Figure 3-3: A barometric loop prevents negative pressure by taking advantage of the vapor pressure of water; unfortunately it requires long stretches of pipe; reproduced from [32] a pull force on the other side of the loop. The height of the loop is specified so that if positively-pressurized water enters the loop, it will not vaporize at the top, but negatively-pressurized water will. This device is robust and extremely simple. Unfortunately, if implemented in an intermittent system, each time the system was turned off, the loop would drain and would only work again when the system provided more than 10m of pressure. Many cities in India do not consistently provide 10m of pressure and therefore this idea too was abandoned. From this brief survey of back-flow prevention devices, houses in India are in need of something as reliable as a vacuum breaker, but that does not entrain air into the system. 3.2 Self-actuated, full-bore pressure regulating valves This thesis sought to design a valve that was self-actuating, full-bore, and pressure regulating. Combinations of two of these three characteristics are reviewed in this section, but no combination of all three was found. Were one to be found, it is likely that it would be extremely complex, unlike the design of the valve described in Chapter 4. As any valve closes, it forces fluid through a smaller and smaller area, increasing 46 Figure 3-4: Self-actuated back-pressure valve with a large diaphragm; reproduced from [34] the local jet velocity and therefore increasing the frictional losses through the valve. A regulating valve is designed to operate in many positions along its closing stroke [33]. A pressure regulating valve is designed to maintain a minimum or maximum pressure either upstream or downstream from the valve. To make efficient use of a regulating valve’s stroke, Skousen’s Valve Handbook recommends that regulating valves be undersized, making even the initial parts of the stroke have a noticeable effect on the frictional losses through the valve [33]. Since many regulating valves follow Skousen’s advice about being undersized, full-bore (i.e. valves that do not reduce the area that the fluid has to pass through) regulating valves are much less common. Still, the most common type of regulating valve is a gate valve [33], which if externally actuated, can be full-bore. Further, an externally-actuated, regulating ball valve could also be full-bore. So while full-bore regulators are rare, they do exist. Self-actuated regulators also exist. Self-actuated gate valves typically have a limited stroke and use a large diaphragm to overcome the frictional forces associated with changing the valve’s position. An example of such a valve is pictured in Figure 3-4, reproduced from [34]; it uses a set of sliding plates with matching orifices to regulate the flow. Self-actuating, full-bore valves are commonly used as steam safety devices, where the full-bore allows the system to release pressure more quickly. One such valve is shown in Figure 3-5. They are frequently actuated with pilot lines, and therefore cannot throttle the flow. 47 Figure 3-5: Self-actuated full-bore valve is pilot operated, designed for emergency relief and not for regulation; reproduced from [35] The combination of all three has not yet been found: self-actuating, full-bore, and regulating. Should such a combination exist, it is likely that it would use a very large diaphragm or complicated pilot lines. To address the problem laid out in Chapter 2, a simple, full-bore, self-actuated, back-pressure regulator is required. The designed valve used a stabilized collapsing tube to accomplish this. 3.3 Collapsible tubes or ‘Starling Resistors’ The design of a self-actuating, full-bore, back-pressure regulating valve was realized with a stabilized collapsing tube. The use of a collapsing tube as a regulator is frequently referred to as a “Starling Resistor” [36] after E.H. Starling, who used one in his experimental apparatus, where the tube only allowed flow through it when the upstream pressure was greater than or equal to the pressure on its outside [37]. Such a tube acts as a self-actuated, full-bore, back-pressure regulator. If subjected to an external pressure of atmospheric pressure, the collapsing tube would prevent negative 48 pressure from passing upstream, and therefore meets the design requirements. Unfortunately, these resistors are known to have an unstable regime in which they oscillate rapidly [38]. No information however was found about methods or mechanisms, theoretical or practical, for removing this instability. While this thesis does not provide the theoretical framework to explain why the design outlined in Chapter 4 stabilized the collapsing tube, a specially-shaped insert placed inside the collapsing tube was found to eliminate much of the instability, preventing it from inducing severe and frequent transients into the system. 49 THIS PAGE INTENTIONALLY LEFT BLANK 50 Chapter 4 The iterative design of a self-actuating, full-bore, back-pressure regulating valve The design of a patent-pending, self-actuating, full-bore, back-pressure regulating valve used an iterative design process. It was developed to increase water quality and distribution equity by eliminating the negative pressures caused by residential booster pumps in India. This chapter outlines the four distinct design cycles that occurred: 1. an initial rapid prototyping and concept generation stage developed the key strategy: a collapsing soft tube; 2. early experiments found the optimal method for stabilizing the collapse of the soft tube: a rigid insert; 3. the alpha prototype took this strategy and packaged it into a product that could be tested in houses in India; and finally, 4. the beta prototype attempted to improve upon two issues encountered during the field trial of the alpha prototype. 51 Table 4.1: Initial functional requirements and design parameters F1 F2 Functional Requirement Design Parameter Keep pressure upstream Back-pressure regulating valve, with a set point from the house positive of atmospheric pressure, i.e. X=0 in Table 4.2 Minimize hydraulic Use a full-bore valve design losses through the device F3 Durable Simple, self-actuated, low part count F4 Inexpensive Simple, self-actuated, low part count Table 4.2: Logic of a back-pressure regulating valve with a set point of X Pressure > X Pressure < X Open Close 4.1 Preliminary concepts and their evaluation The four core functional requirements that drove this design process and the design parameters selected to meet these requirements are detailed in Table 4.1. 4.1.1 Keeping upstream pressure positive To ensure the pressure upstream from the house remained positive, two possible strategies were identified to reduce the efficacy of the booster pump: add a check valve to air that would allow air into the system (entrain air) when it was negatively pressurized, causing two-phase flow; and implement a back-pressure regulating valve that would throttle the flow, inducing cavitation. While the check valve was promising because it would avoid forcing the booster pump to cavitate, it was ultimately abandoned after its inclusion in some of the January 2014 field trial because: 1. The air traveled from the check valve to the booster pump at the velocity of the water and therefore created a significant delay in the control loop, which varied between houses. On houses with a far distance between the valve and the pump, the system would cyclically oscillate (hunt) — instead of finding an 52 equilibrium state. The pump would entrain a pipe full of air before the air reached the pump; when the air did reach the pump, the flow would then stop until the pump had cleared the air; finally, the pump would start again, induce a negative pressure, and the cycle would repeat. 2. The air pathway could double as a contaminant pathway; 3. The air caused consumers’ water to sputter and splash out of the spout; and 4. Some RBPs were elevated and ran on a vacuum above ground level, and air caused these RBPs to lose their prime. Instead of entraining air, the final design throttled the pressure upstream from the pump to reduce its efficacy. The use of pressure as a control variable had less hunting problems because the pressure signal traveled at the speed of sound in the fluid, not at the fluid’s velocity. To use pressure as a control variable, a back-pressure regulating valve was designed. A back-pressure regulating valve has the same logic as a pressure relief valve, whose logic is summarized in Table 4.2. However, unlike a pressure relief valve, whose default state is closed, this back-pressure regulating valve would operate in the open position most of the time. If the valve were placed at the consumer’s house, perhaps next to the water meter, then the set point pressure ‘X’ in Table 4.2 would simply be atmospheric pressure. If instead, the valve were designed to be placed underground and upstream from the consumer’s house, its set point logic would need to be a function of the flow rate to adjust for the variable pressure drop between the valve and the household. While complicated, a regulating valve with a flow-dependent set point has been implemented for large pipes, as in Figure 4-1, reproduced from [39, pp. 266-267]. The underground placement of the valve was ultimately abandoned, however, after feedback from a Tata Center for Technology and Design design-review in August 2013 in Mumbai, which highlighted the relative difficulty of inspecting and maintaining underground equipment. 53 Figure 4-1: Overly complex variable set point valve, reproduced from [39] (b) Fluid is forced through a restriction in (a) Fluid takes two 90 β bends in globe valve, this poppet valve, even while open, reproreproduced from [33] duced from [40] Figure 4-2: Bend and constriction losses in some types of commercial valves 4.1.2 Minimizing hydraulic losses Two major sources of friction occur as fluid passes through a valve: bending and constriction, as in Figure 4-2. For turbulent flow, constriction causes losses proportional to the constricted diameter to the fifth power [13]. Where the cross section is not circular, a modified diameter, called the hydraulic diameter, the cross sectional area and π π·π» , defined in Equation 4.1, is used. π΄ is is the wetted perimeter of the flow path [40, p. 92]. π·π» = 4π΄ π (4.1) The minimum friction therefore comes from a circular cross section with no restriction in diameter, known as a full-bore or full-port design [33]. 54 4.1.3 Concept generation and selection Pursuant of a passively-actuated, low-friction, back-pressure regulating valve with a set point of atmospheric pressure, a structured brainstorm was conducted to generate concepts. Select rough sketches are shown in Figures 4-3 and 4-4. Imposed on these sketches is the reason they were not selected. Key criteria for selecting between the strategies included low hydraulic losses, simple control logic, and low part count. Some promising concepts had large unrestricted flow areas, but large wetted perimeters, and so these designs were eliminated. Ultimately, the full-bore concept of the deflating tube, shown in Figure 4-5, was selected. When the fluid was negatively pressurized, the pressure outside the soft tube caused it to collapse and restrict the flow. When the fluid was positively pressurized, the tube stayed inflated and the fluid saw a full-bore valve, thus fulfilling both functional requirements with only one part. 4.2 Early iterations: searching for stable throttling The first embodiment of the deflating tube concept was simply a very flexible tube attached to barbed fittings and inserted upstream from a booster pump. The tube stayed fully open when positively pressurized, and collapsed closed when negatively pressurized; these two states are depicted in Figure 4-6. A flow test confirmed that the tube prevented the negative pressure downstream from it from traveling to its upstream side, as measured by the pressure recordings shown in Figure 4-7. While the pressure performance of this tube met the functional requirements F1, F2, and F4 from Table 4.1, it had two issues which were thought to reduce the fatigue life — thereby possibly compromising the durability requirement of F3: 1. a large stress concentration at the downstream boundary condition between the flexible tube and the rigid support, marked by the small radius of curvature at the pinch points in the tube in Figure 4-6a; and 2. an unstable regime in which the tube would flutter open and closed, inducing large pressure fluctuations into the system, as can be seen in Figure 4-7. 55 90 degree bend Stable poppet’s are complex Fold and Unfold Pressure and momentum forces hard to balance. 180 degree bend Complex control logic Wetted perimeter large Inflate and Deflate 180 degree bend Momentum influence? Possible Stretch or Pinch How does edge seal? Friction proportional to length less efficient than diameter High part count Figure 4-3: Structured brainstorming sketches, part 1 56 Pull or Collapse Wetted perimeter is large Drag deflection Wetted perimeter is large Pinch points Twist Residual tension in tube is vulnerable to creep Snap Possible Wetted perimeter is large Wetted perimeter is large Figure 4-4: Structured brainstorming sketches, part 2 57 Figure 4-5: Selected concept: a deflating tube as a back pressure valve (a) Negatively pressurized closed tube (b) Positively pressurized open tube Figure 4-6: Detail of fully closed and fully open collapsible tube. Flow is left to right 1 0 −1 Pressure (m) −2 Upstream of Valve Between Valve and Pump −3 −4 −5 −6 −7 −8 −9 70 71 72 73 74 75 Time (s) Figure 4-7: Negative pressure was prevented, but rapid pressure oscillations were observed in the throttled upstream pressure with the unmodified flexible tube 58 Table 4.3: Design variables considered in search of stable operation Variable Effect Pre-buckling the tube Different collapse pattern, but equally unstable Reinforcing the tube Instability removed Smoothing the rigid to flexible transition Equally unstable, but less stress concentrations The supplemental functional requirement that the valve should stably throttle the flow sought to address this second concern. To meet this additional functional requirement, three design parameters were experimented with. The success of each one is summarized in Table 4.3. Three of the prototypes used to gather these results are pictured in Figure 4-8. The reinforcing tube with a gradual transition into the flexible region, shown in Figure 4-8b, proved to both eliminate the instability and to minimize the stress concentrations caused by the flexible tube getting sucked into the downstream end. 4.2.1 Full factorial experiment to determine insert geometry and tube choice Using a full-factorial experiment, the optimal combination of insert size and soft tube material was found to be a silicone rubber tube where the inside diameter of the soft tube matched the cross-sectional perimeter of the insert exactly; the final insert geometry is shown in Figure 4-9. The performance did not depend significantly on the shape of the opening. The parameter variation is summarized in Table 4.4. ‘Insert Size’ refers to the cross-sectional perimeter at the center of the insert (shown in green in Figure 410). It was measured in the percentage that the flexible tube would have to stretch circumferentially to occlude the pipe, calculated assuming no slippage between the insert and the soft tube — later found to be an unrealistic assumption. The over-sized by +40% inserts performed so poorly that they were excluded from the presented results. The undersized (-20%) inserts were prone to fluttering, 59 (a) Pre-buckled design (b) Stiffening support for the inside of the tube (c) Gradual transition from rigid to unsupported Figure 4-8: Three prototypes attempting to remove instability A R0.05 R0.05 A 16 0.53 0.43 SECTION A-A 1.45 (a) CAD render 0.84 0.75 0° (b) Part drawing; units of inches UNLESS OTHERWISE SPECIFIED: DIMENSIONS ARE IN INCHES TOLERANCES: FRACTIONAL ANGULAR: MACH BEND TWO PLACE DECIMAL THREE PLACE DECIMAL Figure 4-9: Optimal insert shape INTERPRET GEOMETRIC TOLERANCING PER: PROPRIETARY AND CONFIDENTIAL THE INFORMATION CONTAINED IN THIS DRAWING IS THE SOLE PROPERTY OF <INSERT COMPANY NAME HERE>. ANY REPRODUCTION IN PART OR AS A WHOLE WITHOUT THE WRITTEN PERMISSION OF <INSERT COMPANY NAME HERE> IS PROHIBITED. MATERIAL SolidWorks Student Edition. For Academic Use Only. NEXT ASSY 60 5 USED ON APPLICATION 4 NAME DATE DRAWN TITLE: CHECKED ENG APPR. MFG APPR. Q.A. COMMENTS: SIZE DWG. NO. REV Old Insert Correct cross section 0.05ra A FINISH 3 SHEET 1 OF 1 SCALE: 1:2 WEIGHT: DO NOT SCALE DRAWING 2 1 Figure 4-10: Optimal insert was one where the cross sectional perimeter matches the inner circumference of the soft tube (0.75inπ = 2.36in) Table 4.4: Insert geometry test matrix; repeated for silicone and latex soft tube Insert Size +40% +20% 0% Linear Opening, as in Figure 4-11 Linear Side Profile, as in Figure 4-12 (a) Top view (b) Side view Figure 4-11: Linear Opening insert geometry (a) Top view (b) Side view Figure 4-12: Linear Side Profile insert geometry 61 -20% Min -4 -5 Linear Open Linear Slope Linear Open Linear Slope Linear Open Linear Slope Total Seconds of Flutter -20 0 20 18 16 14 12 10 8 6 4 2 0 20 Latex Rubber Silicone Rubber Linear Open Linear Slope Linear Open Linear Slope Linear Open Linear Slope -20 0 20 Figure 4-13: Total seconds of flutter for each design parameter Minimum Upstream Pressure (psi) 0 -1 -2 Latex Rubber Silicone Rubber -3 -4 -5 Linear Open Linear Slope Linear Open Linear Slope Linear Open Linear Slope -20 0 20 20 18 16 14 Figure 4-13. The +20% inserts allowed brief, but severe negative pressure shown in Latex Rubber 12 to pass upstream, as shown in Figure 4-14. Finally, silicone rubber was selected Silicone Rubber as 10 the membrane material because it performed better in every category tested. This 8 6 data therefore suggests that a silicone rubber tube should be used with an insert 4 of matching size (0% stretch required). The CAD drawing of this optimal insert is 2 0 shown in Figure 4-9. Linear Open Linear Slope Linear Open Linear Slope Linear Open Linear Slope Total Seconds of Flutter Figure 4-14: Minimum upstream pressure allowed for each design parameter The experimental data did not conclusively suggest one opening style or the other. -20 0 20 It was hypothesized that the linear slope opening (defined in Figure 4-12) would be more resistant to flutter as the membrane stretched over time and therefore it was used instead of the linear opening (defined in Figure 4-11). Should there be a desire 62 to change the opening geometry in the future, there is no reason given here not to. Also of note, the latex rubber proved much more resistant to tearing than the silicone; so should there be a future issue with tearing, the material choice described above should be revisited. Both the alpha and the beta prototypes used the insert shape detailed in Figure 49 and silicone for the soft tube. It should be noted, however, that because of a version tracking issue in the CAD files, the January field trial was conducted with a +20% insert. This mistake was corrected for the March trial. 4.3 Alpha prototype The alpha prototype sought to convert the stabilized tube design of Section 4.2.1 into a product that could be tested in an Indian household. Table 4.5 details the functional requirements created to describe this design objective. F4c-F7 (see Table 4.5) were not requirements in the earliest prototypes, and therefore the design required adjustment. The conclusion of this design process was the product summarized with the bill of materials and exploded assembly view in Figure 4-15 and the 3D CAD model in Figure 4-16. The design featured an insert designed to minimize fatigue in the soft tube, strain transition sections between the soft tube and rigid supports, a supporting wire mesh to carry the positive pressure load, and a sealed outer shell to isolate the device from overly curious customers. The only moving part was the silicone rubber tube of hardness A-35, shown pink in Figure 4-16. When water was positively pressurized, the tube stayed inflated, and did not obstruct the flow, but when water was negatively pressurized, the tube collapsed through the cutaway in the insert (green in Figure 4-16), and throttled the flow. 4.3.1 Durability To minimize the likelihood of fatigue failure, the early prototypes used a gradual transition from the rigid pipe to the partially supported collapsing tube, as in the 63 Table 4.5: Alpha prototype’s functional requirements and design parameters F1 Functional Requirement Design Parameter Sustain positive Back-pressure regulating valve, with of back-pressure atmospheric, i.e. if P if P F2 Minimize frictional losses > < 0, then close more; 0, then open. Use a full-bore valve design through the valve when open F3 F4a Inexpensive Simple and self-actuated, low part count Durable: minimize Simple and self-actuated, low part count complexity F4b Durable: throttle flow Use the stabilizing insert strategy stably F4c F5 Durable: minimize strain Use a gradually opening insert and have mismatches and stress strain transition sections at tube concentrations attachment points Withstand cyclic loading of Soft tube balloons under high pressure, 17m positive pressure and provide support on the outside of the tube full vacuum F6 Be difficult to tamper with No access to tube from outside. No atmospheric vents that could be plugged F7 Safe for drinking water As a first prototype use NSF or FDA contact approvable materials 64 ITEM NO. PART DESCRIPTION This part has NPT threads on both ends and a cutaway of linear slope type Used as strain transitioning material, this short tube is 1/16" thick and has an ID=0.75" QTY. 1 Insert with full radius 2 A-50 Si-Rubber Outer 3 A-35 Si-Rubber Hose This is the valve's actuator. It collapses into the pipe under negative pressures. t=1/16"; ID=0.75" 1 4 5 A-50 Si-Rubber Hose Brass Ferrule t=1/16"; OD=0.75" Crimps the assembly together 2 2 6 PVC Reducer Reduces from 2" to 0.5" unthreaded pipe. McMaster 6826K192 2 7 2" PVC Pipe Nominally 2" schedule 40 1 Constructed of rolled mesh, acts as a positive pressure stress relief for the soft rubber hose (3) 1 8 Mesh tube 7 1 3 2 4 1 2 5 6 8 Figure 4-15: Bill of materials and exploded view of alpha prototype SolidWorks Student Edition. For Academic Use Only. 65 (a) Doubly sectioned view (b) Section view (c) Outer shell Figure 4-16: 3D CAD of alpha prototype 66 SECTION C-C C 1/2" NPT male C R0.05 A R0.05 A B 0.13 SECTION A-A 5.47 6.53 7.28 ° DETAIL D 0.25 SCALE 5 : 1 86 .6 1.98 2.72 3.78 DETAIL B SCALE 2 : 1 9.25 UNLESS OTHERWISE SPECIFIED: NAME DATE Figure 4-17: Alpha prototype insert drawing; units of inches DRAWN DIMENSIONS ARE IN INCHESpart TOLERANCES: FRACTIONAL ANGULAR: MACH BEND TWO PLACE DECIMAL THREE PLACE DECIMAL INTERPRET GEOMETRIC CHECKED TITLE: ENG APPR. MFG APPR. Q.A. PROPRIETARY AND CONFIDENTIAL of Section View C-C TOLERANCING middle segment inPER:the insert’s part drawing in Figure 4-17. COMMENTS: THE INFORMATION CONTAINED IN THIS DRAWING IS THE SOLE PROPERTY OF <INSERT COMPANY NAME HERE>. ANY REPRODUCTION IN PART OR AS A WHOLE WITHOUT THE WRITTEN PERMISSION OF <INSERT COMPANY NAME HERE> IS PROHIBITED. SIZE DWG. NO. REV SolidWorks Student Edition. Another possible location was hypothesized to be at the sealing joint For Academic Usefatigue Only. USED ON NEXT ASSY failure MATERIAL AInsert for thesis FINISH APPLICATION DO NOT SCALE DRAWING SCALE: 1:2 WEIGHT: SHEET 1 OF 1 between 5the rigid support 4and the soft tube, mismatch. The1 soft 2 3 because of a strain tube was made of silicone with Shore A-35 hardness. To ease the transition, a stiffer shore A-50 tube, also of silicone, was used at the interface between the rigid insert and the soft tube, and between the soft tube and the rigid outer crimp. These layers and the assembly stages are detailed in Figure 4-18. To form a durable seal between the insert and the soft tube, a crimp was used to provide a compressive force. For creep resistance, a barbed profile was used on the inside face of the insert. Upon later analysis, however, it was revealed that while the mechanical integrity of a barbed joint under cyclic loading is resistant to creep, the seal — which relies on the compressive force — is not. The barb design is shown in Detail B of the insert’s part drawing in Figure 4-17. To control stress concentrations in the soft tube, a minimum bend radius was incorporated into the insert design, so that even in the fully-closed position, the tube would not need to bend sharply. A fillet radius of 0.05" was used to approximate a 67 Figure 4-18: Alpha prototype’s strain transition layers; far-left the inner strain transition tube is attached, mid-left the flexible tube is stretched over the inner strain transition material, mid-right outer strain transition sleeves are installed, far-right the whole assembly is rolled in wire mesh and crimped 68 fully-rounded edge, as shown in Detail D of Figure 4-17. This choice however would make injection molding the part as a single piece more difficult. Without the inner fillet radius, it could be made using only a center pull. If instead the insert were made as two pieces, then the inner fillet would be insignificant. The necessity of the inner fillet was inconclusively investigated in the fatigue test of Section 4.3.5. 4.3.2 Withstanding system pressure In urban areas in India, drinking water should typically be supplied at 17 meters of pressure, see Table 2.2; in intermittently-supplied areas, this pressure cycles with the system’s intermittency. The valve was therefore required to withstand 17m of cyclic pressure. To meet this requirement additional support for the soft tube was required. This additional support, however, could not compromise the tube’s ease of collapse. Three concepts were investigated to achieve this goal: 1. Create a custom soft tube with a braid inside of it; the braiding material would be easily deformable, but strong under tension. This concept was not pursued because prototyping it would have been extremely expensive and a manufacturer could not be found for a braided tube as soft as A-35. 2. Put the soft tube inside a wire mesh that would fit under the crimped end connections. For prototyping ease, a rolled mesh sheet was used instead of a custom mesh tube. 3. Put the soft tube inside a rigid tube with breathing holes. In this concept, the crimp could be replaced with the rigid tube itself. Figure 4-19 depicts Options 2 and 3. To get the external tube to act as a crimp, a sixjaw chuck was used to crush its ends. Unfortunately, this led to an uneven distribution of compressive forces as might be expected from the shape of the deformed tube seen on the right of Figure 4-19. Under a burst test, Option 2 failed at 48m of water pressure, while Option 3 did not fail at 61m, the maximum pressure of available shop 69 Figure 4-19: Detail of alpha prototype’s two possible external support mechanisms air. Taking burst pressure as a proxy for robustness of the design, the mesh option was pursued. 4.3.3 Tamper resistant To make the design tamper resistant, two risks needed to be eliminated. First, the actuating membrane could be punctured or cut if exposed to the user. Second, the valve’s control logic involved throttling to atmospheric pressure; were this pressure to be calculated using a reference port, for example, plugging the reference port could limit the valve’s functionality. Using a sealed reference-pressure chamber mitigated both of these risks. If the valve had no openings to the outside, puncturing the membrane or plugging the valve would be more difficult. An infinitely large reservoir of gas at atmospheric pressure would allow the system to function as if there were an external port. reservoirs follow the ideal gas law: πΏπ π = Finite-sized πΏπ . The actuated valve was estimated to π have a displaced volume of approximately 3mL, therefore to achieve a <1% error would require at least a 300mL reference chamber. Prioritizing the use of standard components, a 2in PVC pipe was used as an outer shell for the device. It provided a reference volume of 218mL, which led to a 1.3% error, equivalent to the valve having a set point of -0.14m of water pressure. 70 4.3.4 Safe for drinking water contact The final additional requirement for the alpha prototype was that it be safe for drinking water contact. NSF 61 certification is a stringent standard for materials that will be in contact with drinking water. The certification is only mandatory in the US for the testing of lead leeching in devices covered in NSF 61 §9, which applies devices that are "typically installed within the last liter of the distribution system" [41] and that intend to "dispense water for human ingestion" [42, p. 46]. In a house with half-inch piping, this 1L extents about 8m back from the tap, far enough that the valve might have qualified for the 1L range, but as it was not a water dispensing product, compliance was presumed not to be mandatory. To ensure the safety of the valve, efforts were made to source materials that had been used in other NSF-61-compliant products. approvable However, the supply of NSF materials was extremely limited. Instead, all wetted materials used were selected to be made of FDA approvable materials; and none had lead components. The FDA and the NSF only certify products and not materials; therefore it was not possible to use NSF or FDA certified materials, only material that had been certified in other products. For 3D printing or casting, two good options were found: Steralloy FDG Rigid for urethane casting was FDA approvable, and NWRapid’s Nylon 12 for 3D printing had been used in products approved for food-contact by the FDA. Ultimately, the cast steralloy was used because of long-term swelling concerns with a submerged 3D printed part. While the urethane casting process was outsourced, a process photograph is included for completeness in Figure 4-20. 4.3.5 Laboratory fatigue test To maximize the fatigue life of the valve, the effect of the soft tube’s material and the insert’s inner fillet radius were investigated using a full-factorial experiment. The test matrix is outlined in Table 4.6. The first unit failed at 3970 cycles, but because of Christmas holidays the test was 71 Figure 4-20: Process steps in the outsourced casting of the urethane inserts for the alpha prototype Table 4.6: Alpha fatigue test matrix: tube material and inner fillet radius 0.05" inner fillet radius 0.005" inner fillet radius Latex N=3 N=3 Silicone N=3 N=3 72 Figure 4-21: Identifying where fatigued prototypes leaked Mesh seam line Mesh seam line crack Figure 4-22: Detail of failure location and suggested cause in fatigue test not turned off until 20,000 cycles. A post-mortem analysis revealed four prototypes had catastrophically failed; one from each quadrant of the text matrix. Leak locations were identified as in Figure 4-21. In all four failed prototypes, the wire mesh seam aligned with the edge of the stabilizing insert’s opening, creating a pinch point; cracks occurred along this pinch point, as shown in Figure 4-22. Examination of the other prototypes showed some signs of wear, but none of them had the seam alignment found in the four failed prototypes. While the fatigue test did reveal some very useful information about a previously unconsidered variable in the manufacturing of the alpha prototype, it did not answer either question about fillet radius or soft tube material. Since the opportunity for a field trial in January was rapidly approaching, further experimentation was not possible and so the larger fillet radius was used with a silicone tube. 73 Figure 4-23: Alpha prototypes after field testing 4.3.6 Field test The field performance of both the alpha and beta prototypes are combined and discussed in Chapter 6. Seven of the alpha prototypes that were installed during the January 2014 field test are pictured in Figure 4-23. While the pressure performance of these prototypes was very successful, five of eleven of the units failed in shear fracture at the threads. After the first few valve breaks, the remaining valves were installed using a flexible hose connection on the booster pump side of them, as in Figure 4-24. This flexible hose allowed for minor misalignments in the pipe network, but had higher frictional losses through it. 4.3.7 Improvements needed Two improvements were needed after field testing because five of the eleven field units failed in shear fracture at the threads, and several of the units had their reference pressure drift. It was hypothesized that the shear failures were caused by a combination of the cast urethane being brittle and large residual stresses from the misalignment of metal pipes in households. Additionally, Indian plumbers typically only use one pipe wrench and rely on the pipe network to provide the counter force. Together these two large loads applied to a cast material caused frequent shear failure in the prototypes, 74 Figure 4-24: Alpha prototypes’ field test required a flexible hose to reduce the shear load on the valve; from bottom to top, pressure logger, alpha prototype, check valve to air, flexible hose, booster pump as seen in Figure 4-25. The drift in reference pressure was likely caused by a slow leak from the valve into the reference chamber. Once the reference pressure was above atmospheric pressure, the unpressurized valve would stay partially closed when there was atmospheric pressure in the pipe, as in Figure 4-26a. The hypothesized leak pathway is detailed in Figure 4-26b. Several prototypes were observed to have changes in their reference pressure within a few days of being connected to the water system. This slow leak only occurred when there was a large pressure difference between the reference pressure and the pipe pressure. Therefore, the leak initially acted as a one-way leak, allowing only leakage into the reference pressure chamber, not out of it. As such, even a very small would eventually cause a significant pressure drift. The beta prototype sought to maintain the positive field performance of the alpha prototype, while eliminating these two issues. 4.4 Beta prototype While Delhi’s water utility was happy with the field test results of the alpha prototype, they were concerned that the test had only been conducted on older connections 75 Figure 4-25: 55% of alpha field units failed in shear at the threads (a) Several units had the reference set point drift, causing the default position of the valve to be partially closed (b) The hypothesized leak path was under the crimped seal; highlighted in red Figure 4-26: Alpha prototype’s set-point drift because of a slow leak into the reference chamber 76 and focused mostly on continuously supplied connections. To address this concern, a second field test was planned and the beta prototype was built to use in this subsequent test. Additionally, the beta prototype sought to address the two failure modes observed in the alpha prototype: shear failure and drift in the set-point pressure. Addressing these required making compromises in the functional requirements, summarized in Table 4.7. The functional requirement that the valve should be able to withstand axial and radial moments (F4d in Table 4.7) was added to account for the loading forces of plumbers and plumbing. To address the drift in the set-point pressure, a vent was added to the reference chamber, which required a reduction in the tamper-resistance of the valve (F6 in Table 4.7). The following two subsections elaborate on these design decisions and trade-offs. The third subsection, 4.4.3, describes the detailed design of the beta prototype, and finally this section concludes with a summary of the valve’s fatigue and field performance and the next steps for further improvements. 4.4.1 Increasing the valve’s toughness In the alpha prototype, the pipe threads of the cast urethane insert transfered the external load through the threads into the outer PVC shell; this proved not to be strong enough. To maintain the functionality of the valve’s collapse mechanism, but increase the valve’s ability to withstand loading, three options were considered: 1. strengthen the insert, 2. carry the load through the middle reinforcing layer (previously the wire mesh in the alpha prototype), or 3. carry the load through the outer protective shell (previously the PVC tube in the alpha prototype). To simplify the design, separate components of the valve were designed for separate functions. Since the insert’s primary function was stabilizing the collapsing tube, the 77 Table 4.7: Beta prototype’s functional requirements and design parameters F1 Functional Requirement Design Parameter Sustain positive Back-pressure regulating valve, with a set back-pressure point of atmospheric pressure, i.e. if P then close more; if P F2 Minimize frictional losses > < 0, 0, then open. Use a full-bore valve design through the valve when open F3 F4a Inexpensive Simple and self-actuated, low part count Durable: minimize Simple and self-actuated, low part count complexity F4b Durable: throttle flow Use the stabilizing insert strategy stably F4c F4d F5 Durable: minimize strain Use a gradually opening insert and have strain mismatches and stress transition sections at the tube’s attachment concentrations points Durable: able to Separate structure from function; carry the withstand large moments load through the metal outside; no loads on axially and radially the plastic insert Withstand cyclic loading Soft tube balloons under high pressure, of 17m of positive provide support on the outside of the tube pressure and full vacuum F6 Be difficult to tamper Limited direct access from outside; breathing with holes in outer shell and middle shell do not align F7 Safe for drinking water As a first prototype use NSF or FDA contact approvable materials 78 load-bearing requirement was left to other components. This left either Option 2 or 3; the strategy for eliminating set-point drift differentiated between them. 4.4.2 Preventing set-point drift Two design parameters were adopted simultaneously to address the issue of the reference pressure drifting: the strategy of using a sealed reference chamber was abandoned, and the quality of the sealing mechanism was increased. If a sealed reference chamber had been used, the seal would have had to be extremely reliable as even an very slow leak, one milliliter per month for example, would lead to a significant set-point drift over the course of a year, in this example +0.4m. Therefore, it was determined instead that the air chamber should be vented to the atmosphere. This required, however, a sacrifice in the design’s ability to withstand tampering as the air vent could provide access to the soft membrane. Further, if the vent were plugged, the valve’s performance would suffer. One possible future strategy to explore for increasing tamper resistance would be the addition of a large vented chamber such that if the vent were plugged, the chamber’s dead volume would act as a sealed reference chamber and the valve’s functionality would not be significantly compromised. The robust sealing mechanisms of a glued joint and an O-ring were preferred over a crimp. While getting a strong circumferential seal from a crimped joint is possible, as is frequently done with hydraulic hoses, with a very soft tube, assuring the even distribution of the compressive force and accounting for creep was prohibitively challenging. Sealing the soft tube to either the insert, the middle layer, or the outermost layer was required. Both an O-ring seal or an adhesive seal could have been used, but to account for the possibility of creep — which would change the O-ring’s compression — it was determined that the soft tube should be sealed using an adhesive. The only available NSF-certified silicone adhesives had relatively poor peel strength. To minimize the loads applied to the glued joint (using the principle of reciprocity), it was determined that the soft tube should be adhered to either the middle layer or the outermost layer. In this configuration, the larger loads of positive pressure would act 79 Figure 4-27: The peel force was dramatically reduced by turning the membrane insideout at each end of the middle layer to hold the joint sealed, while the smaller negative pressure loads would be applied to the joint. To further strengthen this seal, it was desired to minimize the peel forces acting on the glued joint. To accomplish this, the soft tube was turned inside-out at each end, as shown in Figure 4-27, where the centerline is horizontal and below the figure, the rigid layer is shown in blue, the soft tube in black, and the adhesive in red. In this configuration, when the soft tube is pulled by a negative pressure load down into the flow path (towards the bottom right of Figure 4-27), the glued joint experiences only a shear force. The only mechanism by which a peeling force could be applied to the glued joint would be if the soft tube were loose enough to allow the end sections to balloon out. Even in this scenario however, the surface area of the collapsing section in the center of the valve (not shown in Figure 4-27) would be much larger than the surface area of the potentially ballooning end sections; therefore, the collapsing force would act to limit the ballooning and restore purely shear loading. The inside-out glued-joint design meant that it would be complicated to have the soft tube adhered to the load-bearing member in the structure and therefore the outer layer was taken as the load-bearing element while the soft tube was adhered to the middle layer. In an effort to utilize off-the-shelf components, it was determined that the outer load-bearing shell could be made of standard metal pipe and a simple seal between the outer layer and the middle layer could be accomplished with an O-ring, as in Figure 4-28a, where the black circle represents the O-ring. 80 (a) Sketched sealing mechanism (b) CAD section view (c) CAD doubly sectioned view Figure 4-28: Section views of the beta prototype 81 ITEM NO. PART NUMBER DESCRIPTION QTY. Seamless steel pipe Reducing Coupling 1.25" Seamless Steel pipe; Reamed ID 1.25"NPT female to 0.5"NPT female 3 Stabilizing Insert Cast urethane, linear-side-profile style 1 4 Middle layer Custom turned part; aluminum 1 5 Soft tube A-35 Silicone rubber; t=1/16"; ID=0.75" 1 6 O-ring Viton O-ring Dash 025 2 1 2 1 2 1 2 6 5 4 3 SolidWorks Student Edition. For Academic Use Only. Figure 4-29: Beta prototype’s exploded view 4.4.3 Detailed design of the beta prototype The section and exploded views of the valve assembly are shown in Figures 4-28 and 4-29. The assembly stages of the physical valve are shown in Figure 4-30. The outermost component of the beta prototype was a 1.25 inch seamless, steel pipe (black in Figure 4-28b) with a reamed and sanded inside face, shown in the part drawing in Figure 4-31. Breathing holes were added to the center of this pipe. This pipe connected to two reducers which ended in 1/2" NPT female ends, thus forming the outer and load bearing shell of the valve. Sealing against the reamed and sanded inner surface of the pipe, two O-rings joined the outer shell to a middle supporting layer. The O-rings, grooves, and clearance spacing were designed according to specifications in the Parker O-ring Handbook [43, 82 Figure 4-30: Beta prototype’s assembly stages, left to right: the cast stabilizing insert; the middle layer with the white soft tube glued in place and the O-rings attached; the middle layer inserted into the outer shell; and finally the outer shell with the reducing end-pieces and the set screw installed 4.00 2.00 1.66 1.317 1.25" Male NPT 1.25" Male NPT Figure 4-31: Beta prototype’s outer pipe UNLESS OTHERWISE SPECIFIED: DIMENSIONS ARE IN INCHES TOLERANCES: FRACTIONAL ANGULAR: MACH BEND TWO PLACE DECIMAL THREE PLACE DECIMAL INTERPRET GEOMETRIC TOLERANCING PER: PROPRIETARY AND CONFIDENTIAL THE INFORMATION CONTAINED IN THIS DRAWING IS THE SOLE PROPERTY OF <INSERT COMPANY NAME HERE>. ANY REPRODUCTION IN PART OR AS A WHOLE WITHOUT THE WRITTEN PERMISSION OF <INSERT COMPANY NAME HERE> IS PROHIBITED. MATERIAL SolidWorks Student Edition. For Academic Use Only. NEXT ASSY 5 USED ON APPLICATION 4 NAME DATE DRAWN TITLE: CHECKED 1.25" Seamless Steel pipe; Reamed ID ENG APPR. MFG APPR. Q.A. COMMENTS: SIZE DWG. NO. 83Dimensioned A FINISH SHEET 1 OF 1 SCALE: 1:1 WEIGHT: DO NOT SCALE DRAWING 3 REV Reamed pipe 2 1 0.041 #59 drill spaced at 45 deg +/- 5 deg all around Debur both sides! A 0.246 0.142 DETAIL B SCALE 5 : 1 R0.025 R0.025 Both sides R0.010 ±0.005 SECTION A-A 32 D +0.002 0.050 0.000 Approx R0.005 C B +0.005 0.093 0.000 +0.003 1.307 - 0.003 R0.050 3.950 3.357 2.538 ±0.020 2.288 ±0.020 2.038 ±0.020 1.218 0.625 1.133 R0.050 1.008 Both sides 4.575 4.433 4.329 R0.050 +0.007 0.875 0.000 R0.050 A Both sides 16 DETAIL C SCALE 2 : 1 Both sides 0.593 0.250 DETAIL D SCALE 5 : 1 Both sides Figure 4-32: Beta prototype’s aluminum middle layer; units of inches pp. 4-9] and then the clearance was adjusted according to the pressure and durometer ratings in Oberg et al. [44, p. 2589]. An FDA approved Viton O-ring of size Dash 025 was selected for its resistance to ozone, weather, water, and abrasion. Its hardness was A60. For ease of manufacturing, the middle layer of the beta prototype (gray in Figure 4-28b) was made of aluminum. The mixing of metals in a water pipe network must typically be avoided in order to minimize corrosion, but to quickly develop this prototype, its ease of manufacturing was prioritized over its longevity. The middle layer was made on the CNC lathe and then breathing holes were drilled using a rotary encoder mounted on a CNC milling machine. The complete part drawing is specified in Figure 4-32. Axial translation and rotation of the middle layer with respect to the outer pipe were prevented using a setscrew. 84 The same silicone A-35 tube as in the alpha prototype was used (pink in Figure 4-28b) and bonded to the middle layer using Dow-Corning 732 silicone adhesive. Contact adhesives were considered, but for an underwater bond of two dissimilar materials, their durability would be questionable at best. The 732 was NSF 61 certified and was able to bond aluminum and silicone. The shape of the fold-over was changed between Figure 4-28a and Figure 4-28b so that the elasticity of the soft tube could be leveraged to hold itself in place during the drying of the glue. The innermost component of the beta prototype was the stabilizing insert. The cutaway geometry was the same as that of the alpha prototype, except it was cut short because it no longer needed to interface with the pipe network. No simple method was found in time for March field trial to secure the orientation of the insert with respect to the middle layer. The middle layer’s breathing holes were made axisymmetric so that rotation of the insert would not be problematic. The insert was, however, still free to translate axially. 4.4.4 Laboratory fatigue test The same apparatus as the alpha prototype was used to cyclically test the beta. Because of the small breathing holes in the middle layer of the beta prototype, the test procedure was modified to use asymmetric pressurization cycles of +27m of air pressure on the inside of the valve for 7 seconds, and of +8m of air pressure on the outside of the valve (simulating an internal vacuum pressure) for 5 seconds. A two-by-two full-factorial experiment was done to examine the effect of the insert’s fillet radius and the size of the clearance gap for the soft tube (i.e. the difference in diameters between the insert’s OD and the middle supporting layer’s ID, less twice the thickness of the soft tube); the parameter variation is summarized in Table 4.8. The soft tube’s thickness was 1/16" and therefore the middle layer’s ID=0.875" left no clearance gap. The larger ID was selected so that it could be reamed with a 22.5mm metric reamer. The twelve fatigue tested valves were built according to the design of Section 4.4.3, except the valves did not have a set screw to ensure alignment. 85 After 3000 Table 4.8: Beta fatigue test matrix: clearance for membrane and the insert’s inner fillet radius 0.05" inner fillet 0.005" inner fillet radius radius ID=0.875" (no clearance) N=3 N=3 ID=0.886" (0.011" N=3 N=3 clearance) cycles, one of the middle layers translated axially far enough that the O-ring became un-compressed and leaked. This occurred because the McMaster-Carr reducers used in the fatigue test were longer than the original McMaster-Carr ones measured and designed for (despite having the same part number). To correct this failure mode, spacers were inserted in all of the valves to limit the range of translation and even the “failed” prototype was realigned and testing was resumed. With the spacers, no failures were observed over 26500 cycles of -8m and +27m of pressure on twelve valves. The test was stopped because while new middle and inner layers had been ordered for the March field trial, extra outside shells had not been. Therefore, the true fatigue life was not discovered, it was only determined that the life was greater than 26500. Since no valves failed, the fatigue test did not indicate which clearance gap to use. Instead, since the soft tubes in the beta valves had been qualitatively observed not to collapse as easily as tubes in the alpha prototypes, the larger clearance gap was selected to try to improve the tubes’ ease of collapse. Further, it was hypothesized that breaking in the valves might loosen the membrane and improve this problem. Therefore, six of the twelve newly assembled and unfatigued beta valves were tested for 500 cycles before the March field test. No performance difference between the six un-fatigued and the six fatigued valves was noticed in the subsequent field trial. 86 Figure 4-33: Field modification of the beta prototype 4.4.5 Field test The field performance of the beta prototypes was tested in detail in the March field trial. For this field trial, the spacers used in the fatigue test could not be used because they were not NSF or FDA approved and their metallic content was uncertain. Instead, the valves were fitted with a setscrew through the outside shell that would contact the middle layer. These setscrews were installed at a machine shop in New Delhi, pictured in Figure 4-33. 4.4.6 Improvements needed Without fail, whenever someone in India — engineer, social scientist, or curious bystander — examined the beta prototype, their feedback was that it was “heavy” and therefore was likely to get stolen. This weight arose from the explicit design choice to overdesign the load-bearing structure. An important component of future work would be to better quantify the required strength, so that the valve’s structure could 87 be optimally strengthened. To further reduce the possibility of theft, wherever possible, moving from metal to plastic parts would be desirable as it would minimize the valve’s recycling value. The beta prototype used a set screw to prevent relative motion between the outer shell and the middle reinforcing layer. No such alignment feature was present between the stabilizing insert and the middle layer. After several field tests, it was observed that some of the inserts had migrated to the downstream end of the valve assembly, out of their optimal positioning. A feature to align the insert with respect to the middle layer should be included in future designs. One of the core design requirements of the valve was to prevent negative pressure. To this end the valve’s set point was chosen to be atmospheric pressure. However, as will be explored more thoroughly in Chapter 6, this logic left no room for variation or error in the set point — equivalent to a zero sigma control process where half of the variation would automatically fall outside of specifications. Future prototypes instead could either attempt to have a mildly variable set point (perhaps by changing the axial tension in the membrane), or could intentionally have a set point several standard deviations (sigma) away from the target pressure. Finally, bench-level experiments indicated that the beta prototype’s soft tube did not collapse as easily as the alpha’s. Three hypothesis exist for why this might have been the case: first, the assembly procedure might have left residual axial tension in the membrane; second, the smooth surface area of the middle layer might have been sticking to the membrane; or third, the limited number of small breathing holes in the middle layer might not have provided enough area for the pressure difference to cause a force large enough to create the local deformations in the soft tube required to get larger surface areas exposed to the pressure difference. Since the delayed collapse seemed to be independent of lubrication and constant across prototypes with axially loose membranes, the third hypothesis should be prioritized as the most needed area for improvement in future prototypes. 88 Chapter 5 Experimental and numerical methods 5.1 Experimental methods 5.1.1 Laboratory setup and experimental method A test setup was designed and built to simulate the typical residential booster pump configuration observed in many houses in New Delhi, where the pump is connected directly to the service connection pipe without a sump, as in Figure 2-7. The key features the test setup attempted to imitate were: 1. a static pressure head upstream from the pump, 2. a moderate friction loss between the supply pressure and the pump, 3. a pump that was able to induce negative pressure on its suction side, and 4. frictional losses after the pump. Additionally, for ease of experimentation, it was also required that the setup have the ability to: 1. insert a prototype valve upstream from the pump; 2. measure the pressure before the valve, between the valve and the pump, and after the pump; 89 1” pipe FM 0.5” pipe 0.5” Al. pipe 0.5” Clear pipe RBP Full-bore valve SP P Flow meter Pressure transducer Booster pump Sump pump P P P FM RBP Valve SP Filter Figure 5-1: Laboratory test setup 3. bypass the valve to get baseline data; and, 4. bypass the pump to get baseline data. A schematic of the experimental apparatus which met these requirements is shown in Figure 5-1. The actual setup is pictured in Figure 5-2. The pressure head upstream from the pump was created with an elevated supply reservoir made of a 27x17x15 inch tote bin, filled with water, and placed on top of a 9 foot scaffolding. From there the water flowed out through a port in the bottom of the bin, through a one-inch plastic pipe down to the experimental setup, and into a three-foot straight section of half-inch aluminum pipe. Along this aluminum pipe was an Omega FDT-31 ultrasonic flow meter. The water continued through a throttling gate valve, past a pressure transducer, and into the valve test section where, using full-bore ball valves, the flow was either directed through the valve to be tested, or around it. The parallel sections merged again and the water continued to flow past another pressure transducer. Next, the water flowed through a 1.5 horsepower pump, past a final pressure transducer, and through a section of clear plastic pipe to make cavitation more visible. Finally, the water discharged as a free stream into the air and spilled into the discharge reservoir. 90 Flow Meter Valve Throttling valve Booster pump Discharge reservoir Filter Figure 5-2: Top view of the valve testing area; booster pump is on the right, discharge reservoir is the white tote bin Table 5.1: Laboratory measurement equipment Item Range Accuracy Supplier Ultrasonic Flow Meter 0.5 - 25 GPM reading Omega Discharge-side Pressure 0-100 psia ±0.25% ±0.25% full scale Omega Full vacuum-45 psig ±0.25% full scale Omega Transducers Suction-side Pressure Transducer The discharge reservoir had a submersible pump with a float switch that would automatically pump water to refill the supply reservoir. The set points of the submersible pump were 1.5 inches apart, ensuring the consistency of the pressure head at the top of the experiment. Before reaching the supply reservoir, the water passed through a filter. The accuracy of the measurement equipment is summarized in Table 5.1. The tendency of early prototypes to flutter, documented in Section 4.2.1, was investigated using this test setup. The testing procedure was that for a given upstream resistance, the throttling gate valve on the discharge side of the pump was slowly opened and then slowly closed over the course of three minutes. Then, the valve was throttled back and forth around positions that caused the valve to open or close 91 for another three minutes. All prototypes were tested with high and low upstream resistance. 5.1.2 Fatigue test setup To bound the estimates of valve life, it was desired to force the valves to cyclically experience positive system pressure and then negative booster-pump pressure. After two failed attempts to create an apparatus that used water as the internal fluid, a third test apparatus was successfully built. It cyclically pressurized air on the inside, and then air on the outside of the valve (to simulate negative pressure on the inside of the valve). This apparatus used shop air, a 4-way solenoid, two pressure regulators, and push-to-connect tubes to connect the test units to the air pressure. The experimental schematic for a single valve is shown in Figure 5-3, in which the pictured solenoid control logic was reproduced from [45]. For the actual test, twelve units were connected in parallel and the setup is pictured in Figure 5-4. A limitation of this apparatus design was that if one unit failed catastrophically, air would escape through that unit and prevent the other units from being tested until the failed unit had been removed; the apparatus could therefore only autonomously test until the first unit failed. 5.1.3 Field testing setup and method The goal of the field test was to evaluate the claim that the designed valve prevented booster-pump-induced negative pressure. A Hydreka pressure logger, shown in Figure 5-5, was installed above the ground and immediately upstream from where the valve would be placed. Pressure measurements were made with and without the valve installed. A schematic of the desired layout and its predicted effect on pipe pressure is shown in Figure 5-6. Informed consent from participating houses was verified through the use of an independent translator. Additionally, the valve was installed with an ‘opt-out’ bypass whose use was demonstrated to each participating house; by turning a quarter-turn 92 Solenoid Driven at 1/12 Hz Shop air H=10m R H=17m R Figure 5-3: Vent Layout of the fatigue test; ‘R’ are the two regulators; pressures were typically set to 10m and 17m Figure 5-4: Built fatigue testing apparatus; blue tubes carry 17m pressure for the inside of the valves; pink tubes carry 10m pressure for the outside of the valve 93 Figure 5-5: Pressure loggers from Hydreka, reproduced from [46] valve, the functionality of the valve would be eliminated. The field setup varied slightly between the alpha and beta prototypes because the alpha prototype required a stress-relieving, flexible hose connection. Alpha and beta test setups are shown in Figures 5-7 and 5-8. In the setup for the beta prototype, the control (baseline) data was taken by turning on the opt-out bypass valve, thus measuring the pressure in the pipe without the valve. Because of a strength issue in the alpha prototype, its opt-out bypass had to be made with a flexible hose. Using this bypass as the control was not possible, because its added friction would have skewed the baseline data. Therefore in the alpha prototypes’ test, a separate test unit with the same fittings, except with a straight plastic pipe instead of the valve, was used to gather baseline data; it is pictured in Figure 5-7b. In the design of the field study, it was assumed that the prevalence of negative pressure would vary more between houses, because of booster pump sizes and household demand patterns, than between days at any given house. A crossover study was therefore designed, where the prevalence of negative pressure at a house would be measured when the valve was not installed, and then this baseline would be compared to the prevalence of negative pressure with the valve installed at that same house. The order of control and treatment was varied between houses. In January 2014, Delhi’s water utility, the Delhi Jal Board (DJB), agreed to test prototype valves on 10 houses within their staffing quarters in Pitampura. The staffing quarters were located adjacent to a large reservoir and pumping station, and so the houses in these quarters had exceptionally good pressure. Only apartments three or more stories above the ground used booster pumps. Even among these apartments, 94 BP Supply Pressure M V Pressure P = Booster Pump = Water Meter = Valve =Pressure Logger Tank 2 Story Home P V Ground Level Distribution Main M BP P<0 Service Connection Pipe Figure 5-6: Field test schematic and predicted pressure performance Stress Relieving Hose Stress Relieving Hose Booster Pump Control Pipe Valve Pressure Sampling Point Bypass Pressure Logger Pressure Logger (a) Alpha prototype during installation (b) Control measurement for alpha prototype Figure 5-7: Three prototypes attempting to remove instability 95 Pipe to overhead tank Booster Pump Valve Pressure Logger Bypass is open to get control measurement DJB Supply Pipe Pressure Sampling Point Figure 5-8: Field equipment arrangement for beta prototype; currently taking control measurements — valve is bypassed negative pressure lasting more than a few seconds was only observed at one service connection. valve. As such, these quarters were not an effective testing location for the Two more connections were identified by the DJB in Pitampura which had more representative supply pressures. To supplement these data points, a convenience sample of houses in Azad Market was taken. neighbors. All five Azad Market houses were The final January connection was also a convenience sample and the house was that of a potentially interested industrial partner’s home. The quantitative characteristics of these houses are summarized in Table 5.2. After reviewing the gathered data, the DJB was concerned that the data was not representative of Delhi because it did not sample enough intermittently connected houses. Further, the DJB wanted to know if the baseline problem of negative pressure was present only on old connections with high frictional losses (like those sampled), or if it would also be a problem on newly connected houses. The latter question is of particular importance because were the valve to be purchased by a water utility it would likely be installed as part of a service connection replacement project. However, would that service connection replacement negate the valve’s importance? To address both of these concerns, a second round of field trials was done in March 2014 with the beta prototype. Finding new connections meant working with some 96 Table 5.2: House connection characteristics, where high pressure is >6m, and medium pressure is between 2 & 6m, and low pressure is between 0.5 & 2m; C/I indicates if the connection was Continuous or Intermittent; O/N indicates if the connection was Older than 2 years or Newer than 1 year Data Label Location High Medium Low C N Pressure Pressure Pressure or or (hrs/day) (hrs/day) (hrs/day) I O J01 Azad Market 5 16.5 1.9 C O J02 Azad Market 4.6 17.3 0.5 C O J03 Azad Market 4.7 17.1 0.6 C O J04 Azad Market 4.7 18.8 0.3 C O J05 Azad Market 0 12.9 9.9 C O J06 Pitampura 10.8 10.1 1 C O J07 Pitampura 4 11.7 4.1 C O J08 Vivek Vihar 1.1 2.9 0.7 I O J09 Pitampura 23.4 0.1 0.1 C O J10 Pitampura 23.9 0 0 C O J11 Pitampura 23.8 0 0.1 C O J12 Pitampura 23.9 0 0 C O M01 Vasant Vihar 0.9 7.6 7.4 C N M02 Vasant Vihar 0 0.3 3.3 I N M03 Shivalik 0 0.3 3.3 I N M04 Shivalik 2 0.4 4.2 I N M05 Lado Sarai 0.4 0.4 2 I O M06 Lado Sarai 0.3 0.4 0.8 I O M07 Lado Sarai 0.7 0.3 2.1 I O of the private water companies in New Delhi, which had recently been upgrading service connections. A total of four new house connections were included in the trial. The private companies selected which houses would be included in the study based on the request that houses be selected where: booster pumps were used, system pressure was positive during supply hours, and where new house connections had been recently made. An additional three houses were included in the author’s plumber’s neighborhood. 97 5.2 Hydraulic notation and units In the field of municipal water systems, one often accounts for pressure, velocity, and frictional losses in units of head, and π π», is defined as pressure head, velocity head, π» = π where ππ π is the pressure, and π head losses. Pressure is the density of water, is the gravitational acceleration. Similarly, the velocity head is π2 , where 2ππ π is the velocity [47, p.8]. In this thesis, unless otherwise specified, pressure will refer to pressure head and will be measured in units of meters of water. Additionally, the pressures specified herein will always be given in gauge pressure where 0m is atmospheric pressure. One commonly used empirical equation for determining frictional losses π»loss in meters, is the Hazen-Williams equation [47, p.14]: π»loss = where π 10.67πΏπ π1.852 1.852 4.87 πΆπ»π π· 3 is the volumetric flow rate [π /π ], length of pipe [m], and πΆπ»π π· is the pipe diameter [m], (5.1) πΏπ is the is a tabulated empirical constant for a given pipe material and age. Another is the Darcy-Weisbach equation [47, p.10]: π»loss = π where π 8πΏπ2 π 2 π·5 π (5.2) is the friction factor, independent of Reynolds number for wholly turbulent flow. 5.3 Quantifying the intrusion risk Contaminant intrusion describes the phenomenon of a contaminating substance entering into the potable water supply. A model to quantify the relative risk of contaminant intrusion was required so that the risks associated with different pipe-pressure histories could be compared. Such a method could, for example, determine which scenario had the higher risk of contaminant intrusion: four hours where a water system was 98 unpressurized and its pipes were empty, or 15 minutes where the same system was filled with water that had a pressure of -3m. This section develops such a method. Contaminant intrusion in a water distribution system can only occur when three requirements are met: when there is a pathway for contamination, when contaminants are present, and when there is a pressure gradient that allows contaminants into the pipe. A perfect contaminant intrusion model would consider all three of these requirements and their interdependence in a given distribution system. In continuously operated systems, negative pressure occurs only in extreme circumstances or during transient events. Modelers of such systems, therefore, typically focus on carefully quantifying the location, duration, and magnitude of these infrequent occurrences of negative pressure. They then approximate the size of the pathway for contamination by the physical leakage rates, and finally, they conservatively assume that contaminants are omnipresent [48, 29]. In intermittently operated systems, an alternative modeling approach was proposed by Vairavamoorthy et al. [29]. They argued that in intermittent water systems, because of the long stretches with no pressure, intrusion is not pressure limited, but limited by the presence of contaminants and contamination pathways. As such, their contaminant intrusion model focused on mapping and predicting the presence and overlap of contaminants and contamination pathways. Applying their method to a city in India which had only one hour of water supply per day, they created a 3D spacial model for the relative location of every water pipe in the system with respect to every possible contamination source, as in Figure 5-9, reproduced from [29]. In areas where water pipes were close to contamination sources, Vairavamoorthy et al. then used coupled models for pipe condition to assess the likelihood of a contamination pathway and models for contaminant flow through the soil around the contamination sources. Together these coupled models assessed the likelihood of contaminant intrusion. The advantage of Vairavamoorthy et al.’s method was that it identified the highest risk areas in the existing system that should be addressed. The recommended course of action was therefore very accessible to the partner water utility as it focused on 99 Figure 5-9: The spacial method used by Vairavamoorthy et al. to model where along a distribution system contaminants would be present, reproduced from [29] incremental improvements such as replacing certain sections of pipe and upgrading certain sections of sewer. Unfortunately, the model was extremely information in- tensive. It relied not only on 3D GIS information about the location and depth of water pipes, sewers, and standing bodies of water, but it also required estimates of the material and age of these water pipes. Further, the model neglected the intrusion potential of saturated soil during or after rains, found to have measurably reduce water quality in a different Indian city [4]. Vairavamoorthy et al.’s model, while theoretically interesting, could not be applied to the data considered later in this thesis for the same reason it has not been used by most Indian utilities: it requires information that is not readily available to utilities and which would be very expensive to gather. Therefore, in this thesis, the more conventional approach to modeling intrusion, which assumes omnipresent contaminants, was used. 5.3.1 Water leakage and its pressure dependency Historically, leaks of water out of pipes were modeled as a free jet leaving a sharp edged orifice with the equation [49, 50]: π = where π √οΈ was the exit velocity of the jet and 2ππ» π» (5.3) the pipe pressure. To convert this to a volumetric leakage rate, the orifice area was assumed to be constant and the leakage 100 flow rate, ππΏ , was found to be [9, p.101]: ππΏ = π΄eq where π΄eq √οΈ 2ππ» (5.4) was an equivalent area of leakage which summed all of the leaks in the system. It was therefore common practice to assume that the leakage flow rate would scale as √ π» and to lump the equivalent leak area under the empirical coefficient πΏπΆ [48]: π πΏ = πΏπΆ × where π»π √οΈ π»π for π»π > 0 (5.5) was the pressure in the pipe. The industry standard changed, however, after it was shown that the pipe’s pressure affected the size and shape of the leaking orifice; for example, a higher pressure may act to widen cracks. To account for this effect, the American Water Works Association records that common practice is to use a coefficient of about one: π∝π» [9]. 5.3.2 Quantifying the potential for contaminant intrusion This section uses the traditional method for modeling contaminant intrusion in continuously supplied water systems; intrusion is modeled as pressure and pathway dependent and contaminants are assumed to be omnipresent. To account for pathway dependence, this section builds on the methodology of Ebacher et al. [48], which used a pipe’s leakage rate as a proxy for the size of that pipe’s contamination pathway. In the historical model for clean water leaking out of a pipe, the leakage coefficient, πΏπΆ , was used to account for the size of the available pathway for leakage, as in Equation 5.5. Contaminant intrusion can be modeled as a leak flowing in the opposite direction; instead of a potable water jet exiting through a crack, a jet of sewage might enter through that same crack. This section develops parallel sets of equations for leakage rates and contaminant-intrusion rates. If pressure had no effect on the size of the leakage area, then the leakage coefficient 101 πΏπΆ would be shared between these parallel equations. The similarities between leakage and intrusion equations, however, breakdown when the pressure-dependent size of the leakage area is considered. For example, increasing the magnitude of positive pressure in a pipe will cause leaks to flow out of the pipe faster and will increase the size of some of the cracks; similarly, increasing the magnitude of negative pressure in a pipe will cause contaminants to flow into the pipe faster, but it will not increase the size of any of the cracks and might act to close them. Quantifying the difference between these two scenarios was beyond the scope of this thesis and so the pressure dependent size of orifices was neglected, allowing the use of the historical leakage equation (Equation 5.5). This simplifying assumption was also made by Ebacher et al. [48]. Taking and π»π ππΌ as the volumetric flow rate of contaminants flowing into the pipe, π»π as the respective pressures of water inside the pipe and of contaminants outside the pipe, and using the same leakage coefficient πΏπΆ as in Equation 5.5, we find Equation 5.6: β§ β¨ πΏ × √οΈπ» − π» : (π» − π» ) > 0 πΆ π π π π ππΌ = β© 0 : (π»π − π»π ) ≤ 0 (5.6) The pressure of contaminants is typically less than, or equal to, their depth underground. In the field setup described in Section 5.1.3, the reference for the pipe’s pressure, π»π , was taken at or just above ground level. In this pressure reference plane, the pressure of contaminates at the underground pipe minus the depth of the pipe is π»π . Since π»π will not typically be above zero, it is taken to be zero and neglected from the equation. This assumption overestimates the amount of contamination. Adapting Ebacher et al.’s methodology to account for temporal and spacial variation in pressure, a daily averaged intrusion volume of pipe of length time, π‘, πΏπ , and observation window π₯, integral of Equation 5.6: and space, 102 π ππΌ (defined over a defined section in days) was found by taking the Where ∫οΈ π ∫οΈ √οΈ πΏπΆ ππΌ = (−π»π* (π‘, π₯))ππ₯ππ‘ π πΏπ 0 β§ β¨ π» (π‘, π₯) : π» (π‘, π₯) < 0 π π π»π* (π‘, π₯) = β© 0 : π» (π‘, π₯) ≥ 0 (5.7) (5.8) π Similarly, taking the time integral for the daily averaged outward leakage rate, ππΏ , but assuming constant pressure along the length of the pipe, allows Equation 5.5 to be rearranged to : Where ∫οΈ πΏπΆ π √οΈ ** ππΏ = (π»π (π‘))ππ‘ π 0 β§ β¨ π» (π‘) : π» (π‘) > 0 π π π»π** (π‘) = β© 0 : π»π (π‘) ≤ 0 (5.9) Flow-rate changes at a single house in isolation do not affect the pressure in the distribution main because head loss scales as pipe diameter to approximately the fifth power, as in Equations 5.1 and 5.2. When a booster pump turns on and increases the flow rate in the service connection pipe, the pressure drops from the system pressure in the distribution main, booster pump π»π΅π (π‘). π»π (π‘), to the measured pressure just upstream from the This leads to the pressure profile shown in Figure 5-10, where the horizontal axis is the effective length of the pipe between the distribution main and the suction inlet of the RBP. All local losses, e.g. elbows, have been converted to their effective length of straight pipe. To proceed with the integration in Equation 5.7, the simplifying assumption was made that the leakage coefficient was evenly distributed along the effective length of the service connection. Figure 5-10 defines point π as the x-coordinate of zero pressure and 103 πΏπ as the length of the pipe. Using Figure 5-10: Pressure drop along the service connection, from the distribution main to the suction inlet of the RBP the definition of π»π* (π‘, π₯) in Equation 5.8, yields: β§ β¨ π» (π‘, π₯) : π₯ > π π π»π * (π‘, π₯) = β© 0 :π₯≤π (5.10) Combining Equations 5.7 and 5.10, πΏπΆ ππΌ = π πΏπ ∫οΈ 0 π ∫οΈ πΏπ √οΈ (−π»π (π‘, π₯))ππ₯ππ‘ π 104 Linearizing π»π (π‘, π₯) based on Figure 5-10 * π»π (π‘, π₯) = π»π (π‘) − πΏπ₯π (π»π (π‘) − π»π΅π (π‘)) √οΈ ∫οΈ ∫οΈ π πΏ * (π‘))ππ₯ππ‘ ∴ ππΌ = ππΏπΏπΆπ 0 π π −(π»π (π‘) − πΏπ₯π (π»π (π‘) − π»π΅π β§ β¨ π» (π‘) : π» (π‘) < 0 π΅π π΅π * Where π»π΅π (π‘) = β© 0 : π»π΅π (π‘) ≥ 0 [οΈ ]οΈπΏπ * (π‘) π»π (π‘)−π»π΅π 3/2 ∫οΈ 2( π₯−π» (π‘)) π π πΏπ ππ‘ ∴ ππΌ = ππΏπΏπΆπ 0 π»π (π‘)−π» * (π‘) π΅π 3( ) πΏπ π But from Figure 5-10 π= β§ β¨ π»π (π‘)πΏπ * (π‘) π»π (π‘)−π»π΅π : π»π (π‘) ≤ 0 β© 0 2πΏπΆ ∴ ππΌ = 3π : π»π (π‘) > 0 ∫οΈ 0 π β§ βͺ βͺ βͺ β¨ * (π‘))3/2 (−π»π΅π * (π‘) π»π (π‘)−π»π΅π * (−π»π΅π (π‘))3/2 −(−π»π (π‘))3/2 * (π‘) π»π (π‘)−π»π΅π βͺ βͺ βͺ β© √οΈ−π» * (π‘) π΅π β« βͺ : π»π (π‘) > 0&π»π (π‘) ΜΈ= π»π΅π (π‘) βͺ βͺ β¬ : π»π (π‘) ≤ 0&π»π (π‘) ΜΈ= π»π΅π (π‘) ππ‘ βͺ βͺ βͺ : π» (π‘) ≤ 0&π» (π‘) = π» (π‘) β π π π΅π (5.11) To better visualize the connection between the pressure history and the intrusion and leakage rates, the un-normalized forms of Equations 5.9 and 5.11 were used in Section 6.2. These un-normalized forms were defined as the relative total volume of intrusion and leakage, ππΌ′ (π‘) and ππΏ′ (π‘), ππΏ′ (π‘) ππΏ (π‘) = = πΏπΆ as calculated in Equations 5.12 and 5.13. ∫οΈ π‘ √οΈ (π»π** (π‘))ππ‘ β§0 β¨ π» (π‘) : π» (π‘) > 0 π π ** π»π (π‘) = β© 0 : π» (π‘) ≤ 0 π 105 (5.12) β§ * (π‘))3/2 βͺ (−π»π΅π : π»π (π‘) > 0&π»π (π‘) ΜΈ= π»π΅π (π‘) * (π‘) βͺ π» (π‘)−π» ∫οΈ π‘ βͺ β¨ π * π΅π π (π‘) 2 πΌ (−π»π΅π (π‘))3/2 −(−π»π (π‘))3/2 ππΌ′ (π‘) = = : π»π (π‘) ≤ 0&π»π (π‘) ΜΈ= π»π΅π (π‘) * (π‘) βͺ πΏπΆ 3 0 βͺ √οΈ π»π (π‘)−π»π΅π βͺ β© −π» * (π‘) : π»π (π‘) ≤ 0&π»π (π‘) = π»π΅π (π‘) π΅π β« βͺ βͺ βͺ β¬ ππ‘ βͺ βͺ βͺ β (5.13) 5.3.3 Defining the contaminant intrusion ratio The Contaminant Intrusion Ratio (CIR), is here defined as the ratio of the volume ππΌ , of potentially intruded contaminants, to the volume of clean water leakage, Over a common time interval, the ratio of volumes of averaged intrusion and leakage rates ππΌ and ππΏ ππΌ and ππΏ ππΏ . is the same as the ratio defined in Equations 5.9 and 5.11. The CIR is therefore defined as: 2 3π ′ CIR = ππΌ ππΌ = = ππΏ ππΏ β§ * (π‘))3/2 (−π»π΅π βͺ βͺ : π»π (π‘) > 0 * (π‘) βͺ π»π (π‘)−π»π΅π ∫οΈ π ′ β¨ (−π» * (π‘))3/2 −(−π» (π‘))3/2 π π΅π : π»π (π‘) ≤ 0 * (π‘) 0 βͺ π»π (π‘)−π»π΅π βͺ √οΈ βͺ β© −π» * (π‘) : π»π (π‘) = π»π΅π (π‘) π΅π √οΈ ∫οΈ ππ 1 π»π** (π‘)ππ‘ 0 π β« βͺ βͺ βͺ β¬ ππ‘ βͺ βͺ βͺ β π (5.14) β§ β¨ 0 : π»π΅π (π‘) > 0 * where π»π΅π (π‘) = β© π» (π‘) : π» (π‘) ≤ 0 π΅π π΅π β§ β¨ π» (π‘) : π» (π‘) > 0 π π ** and where π»π (π‘) = β© 0 : π» (π‘) ≤ 0 π Care must be taken in the definitions of the averaging periods, denominator and numerator of Equation 5.14. ππ ππ and π ′, in the is taken as the total duration of the control and treatment tests at a particular connection so that the leakage rate will be the average outward leakage rate of both the control and treatment observations. Since the ultimate goal was to compare the CIR with and without the valve, recorded pressures during field studies were substituted for 106 π»π΅π (π‘). To calculate the baseline CIR of a customer’s connection, the numerator of Equation 5.14 used control data from that connection and therefore π′ was the duration of the control measurement. To calculate the treatment CIR, treatment pressure data was used in the numerator and π′ was the duration of the treatment study at that house. The assumptions made in the derivation of Equation 5.14 are listed in Table 5.3; their effects are summarized in Table 5.4. 5.3.4 Contaminant concentration at the consumer’s tap The Contaminant Intrusion Ratio (CIR) is the ratio of the volume of potentially intruded contaminants to the volume of water leakage on that same pipe over the same time interval. Where the physical losses from leakage are known as a percentage of clean water delivered, can be found. πΏπΏ, the ratio of intruded contaminants to clean water delivered This ratio will be referred to as the Intrusion Concentration (IC), defined by Equation 5.15: πΌπΆ = πΆπΌπ × (πΏπΏ)along the service connection (5.15) This thesis focuses specifically on the contamination occurring along the service connection pipe. Therefore to calculate the IC, the leakage rate along this pipe specifically must be estimated. The American Water Works Association’s manual on water loss states that the majority of real losses (see Table 2.1 for the definition) occur along household connections [9, p.101]. The Delhi Jal Board estimated real losses of 20-25% throughout their system [11]. Combining these two numbers resulted in an estimated 10% real losses along the service connection. As an example, consider a customer connection measured to have a CIR of 5%. Assuming a 10% leakage level along this pipe, the final water delivered to the customer would have an IC =0.5%; their drinking water could be 0.5% rainwater, or even sewage! While qualitatively, this number is much higher than desirable, the quantitative significance of this metric is still unclear. The following section considers how 107 Table 5.3: The assumptions behind Equation 5.14 Assumption Assumption Number 1 Contaminants are always present in the vicinity of the pipe 2 Area for leakage out of the pipe does not increase with pressure 3 Area for intrusion into the pipe does not decrease with larger magnitude negative pressure 4 Areas for leakage and intrusion are the same 5 Pressure of contaminants is the same as their depth underground 6 Driving pressure for leaks is constant along the service connection 7 Pressure in the distribution main is not dependent on the service pipe connection flow rate 8 Leakage and intrusion rates are evenly distributed along the service connection’s effective pipe length Table 5.4: The effects of Equation 5.14’s assumptions Assumption Effect on calculated Effect on calculated leakage rate intrusion rate 1 Greatly overestimates 2 Underestimates 3 Overestimates overestimates Overestimates Unknown 5 Usually Usually overestimates overestimates significantly significantly Overestimates 7 8 Greatly Overestimates 4 6 Effect on CIR Underestimates Slightly Slightly underestimates underestimates Varies, but mean Varies, but mean Varies, but mean effect is zero effect is zero effect is zero 108 this metric relates to human health risks. 5.3.5 Connecting contaminant intrusion to health risks A more common metric for water contamination in distribution systems is the concentration of indicator organisms, either total coliforms or E. coli, measured in colony forming units (CFU) per 100mL. The two most likely sources of contamination are sewage and rain or surface water. Assuming the potable water coming into the service connection has zero CFU and further assuming perfect mixing such that the intrusion is spaced out over the whole supply time yields the following expected water quality at the customer’s connection: πΆπΉ ππ»π» = πΌπΆ × πΆπΉ πcontaminant (5.16) = πΏπΏ × πΆπΌπ × πΆπΉ πcontaminant The contamination concentration at the household, πΆπΉ ππ»π» , is therefore linearly dependent on the contaminant concentration in the intruding volume, πΆπΉ πcontaminant . No information was found regarding the contamination of rain water in soils; instead, contamination levels in surface water was used as an estimator. Sargaonkar and Deshpande classified excellent, acceptable, and slightly polluted surface water in India as having a total coliform count of less than 50, 500, and 5000 CFU/100mL, respectively [51]. Quantifying sewage contaminants, Miescier found contamination levels flowing into nine municipal waste water facilities ranged from: total coliforms 106 to 108 CFU/100mL and E.coli 105 to 108 CFU/100mL. The World Health Organization’s guidelines for surveillance and control of community water supplies give an example risk categorization for water supplies based on E. coli concentration in drinking water, shown in Table 5.5 [52, p.78]. Connecting these risks to the calculated CIR values requires four assumptions, summarized in Table 5.6. Notably, these assumptions conservatively overestimate the health risk associated with negative pressure because they represent a worst case scenario. These 109 Table 5.5: WHO’s example thresholds for health risks from E. coli contamination, from [52, p.78] Concentration of E. coli CFU/100mL Risk Level 0 Safe: meets requirements 1-10 Low 10-100 Intermediate 100-1000 High >1000 Very high Table 5.6: Assumptions required to connect CIR values to health risks Assumption Health risk level Improvement in health risk Water entering the service connection Underestimates Overestimates Overestimates Overestimates Overestimates Overestimates The physical leakage level along the Unknown Unknown service connection pipe is 10% accuracy; linear accuracy; linear effect effect pipe is free of contamination The entire length of the service connection pipe is exposed to contamination The contaminant is sewage with E. 5 concentration of 10 CFU/100mL coli assumptions allow Table 5.5 to be rearranged to provide maximum allowable CIR levels, as in Table 5.7. Notably, the risk levels increase with the order of magnitude of contaminant concentration. For intrusion reduction to have a significant health impact therefore, it must reduce the CIR value by an order of magnitude. 5.4 Other numerical methods 5.4.1 Monte Carlo sensitivity study A Monte Carlo simulation with N=1000 was conducted to study the sensitivity of the results of Section 6 to variation in the baseline pressure history of the connections, 110 5 Table 5.7: Allowable IC and CIR levels for sewage (10 CFU/100mL) as the contaminant; assumptions are detailed in Table 5.6 Targeted Risk Level Maximum IC Maximum CIR @ 10% leakage Safe: meets requirements <0.001% <0.01% Low 0.001-0.01% 0.01-0.1% Intermediate 0.01-0.1% 0.1-1% High 0.1-1% 1-10% Very high >1% >10% sensor error, and incorrect assumptions about contaminant pressure. A normally distributed, zero-mean, mean shift was applied to each half of the crossover data. I.e. for a specific house, the control and the treatment pressure histories each received a different random error. The variance in the simulated error was determined by the precision of the pressure loggers measured in Appendix A and was taken to have an approximate standard deviation of π= 0.45π . To summarize the range of possible values determined by the 3 simulation, x and y error bars that spanned, unless otherwise noted, the 1st and 99th percentile were included with the data. 5.4.2 Determining upstream system pressure during booster pump events The method for quantifying the intrusion risk caused by negative pressure developed in Section 5.3 requires that the upstream system pressure be known. During field trials, only one pressure measurement was taken and so an algorithm was developed to estimate the upstream pressure. It identified sudden pressure drops caused by booster pumps and stored the pressure before the pump turned on as the upstream system. The upstream pressure was then assumed to be constant until a sudden pressure rise was observed, signaling the turning off of the booster pump. A crude threshold and delay algorithm was used; if the pressure changed by more than 0.75m over two minutes, a pump start or a pump stop was identified. Further, the maximum 111 J09 Upstream Pressure Estimate Upstream Estimate Control Upstream Estimate Valve M01 Upstream Pressure Estimate 12 4 10 3 8 Pressure (m) Pressure (m) 2 6 1 0 4 −1 2 −2 0 01/18/14−09:30 09:40 Date 09:50 04/06/14−15:00 18:00 21:00 04/07−00:00 Date Figure 5-11: Two examples of upstream pressure estimation; the algorithm worked well for cases, like M01, where the pressure drop and rise caused by a booster pump was abrupt, but worked less well for cases, like J09, where the action of a booster pump was less clear pump action was assumed to be 5 hours. Two examples of the algorithm’s mixed accuracy are shown in Figure 5-11. Where the upstream estimate prematurely declines, as in J09, the contamination risk at that house is overestimated. In future work, finding a more robust method of identifying booster pump cycles would make the contamination estimate more accurate. 112 Table 5.8: Measured supply times House Identifier First Supply Time Second Supply Time J08 5:50AM - 9:05AM 6:55PM - 9:05PM M02 2:50AM - 11:25AM 4:10PM - 11:35PM M03 3:45AM - 7:45AM 3:45PM - 7:05PM M04 3:40AM - 7:40AM 3:50PM - 7:05PM M05 2AM - 4:35AM 5:15AM - 6:40AM M06 2:05AM - 4:35AM 5:15AM - 6:30AM M07 2AM - 4:40AM 5:05AM - 6:35AM 5.4.3 Determining the supply hours of intermittently supplied houses To determine the valve’s impact on water quality during supply hours, the supply hours for each intermittently supplied house were measured and are summarized in Table 5.8. The measurement was conducted by overlaying the control and treatment data from each house and visually selecting the two periods in the day where the most pressure variation occurred. Three examples of this are shown in Figure 5-12. 113 Pressure History Overlaid J08 10 Pressure (m) Control Valve 5 0 −5 01/00/00−00:00 03:00 06:00 09:00 12:00 Time 15:00 18:00 21:00 01/01−00:00 Pressure History Overlaid M02 Pressure (m) 15 Control Valve 10 5 X: 11:36 PM Y: 0.2038 0 −5 01/00/00−00:00 03:00 06:00 09:00 12:00 Time 15:00 18:00 21:00 01/01−00:00 Pressure History Overlaid M07 20 Control Valve Pressure (m) 15 10 5 0 −5 01/00/00−00:00 03:00 Figure 5-12: 06:00 09:00 12:00 Time 15:00 18:00 21:00 01/01−00:00 Supply times were determined manually for each intermittently con- nected house; results are summarized in Table 5.8 114 Chapter 6 Results The key design objective of the valve designed in Chapter 4 was to limit the ability of booster pumps to create negative pressure in the service connection pipe, thereby decreasing contamination and increasing distribution equality. This chapter evaluates the valve’s performance in two rounds of a crossover study conducted in New Delhi in January and March of 2014. In this study, the pressure upstream from a particular house was observed for at least 24 hours both with and without the valve at a sampling frequency of 0.1Hz. Because of the limited scope of this trial, it was not possible to evaluate the valve’s impact on distribution equity. Instead, three metrics were used to measure the valve’s ability to reduce negative pressure and contamination risk; 1. Negative pressure duration: the average number of minutes per day a customer had negative pressure in their service connection pipe; 2. Contaminant Intrusion Ratio (CIR): the volume ratio of possibly intruded contaminants to leaked water in the customer’s service connection pipe (more formally defined in Section 5.3.3); and, 3. Health risk level: the World Health Organization’s example health risk level associated with a given CIR value — the assumptions behind this conversion are outlined in Section 5.3.5. 115 Before comprehensively evaluating the valve’s performance, this chapter first examines the valve’s impact on booster pump cycles in Section 6.1. Second, the performance metrics are introduced by means of a case study. Third, the sources of uncertainty are introduced with a second case study. Next, in Sections 6.4 through 6.7, the valve’s effect on these metrics is shown to vary significantly between different types of customer connections. The valve’s performance is evaluated using the four categories: 1. overall performance; 2. performance on continuously-supplied houses; 3. performance on intermittently-supplied houses, particularly while the system is operating; and, 4. overall performance while the system is operating. Finally, the chapter ends with a discussion about how the valve’s effect on distribution equality could be tested in future work. The results in this chapter demonstrate that: β the valve was most effective at limiting severe negative pressure, preventing 96% of all pressure less than -1 meter; β the signal-to-noise ratio in the data gathered from intermittently-connected houses was low; β the valve reduced the duration of negative pressure by an average of 53 minutes per day, per connection while the system was operating; β the valve reduced the risk of contamination by a median of 80% while the system was operating; β the negative pressure caused by booster pumps poses a significant health risk while the system was operating; and 116 β the valve not only took seven of the fourteen houses out of the very high high safe health risk levels, but also brought six of those seven into the and category. 6.1 Effect of the valve on a single pump cycle The valve’s effect on customers’ residential booster pumps (RBPs) depended on the water supply pressure. Figure 6-1 shows examples of single OFF-ON-OFF cycles for RBPs installed on houses with high, medium, and low pressure. On most tested highpressure connections (above 6m of pressure), when a RBP was turned on, no negative pressure was induced, as shown in the top panels of Figure 6-1. The valve was neither supposed to nor did have any effect in this scenario. Conversely, on medium-pressure connections (pressure between 2 & 6m), RBPs did induce negative pressure. In these medium-pressure cases, the valve effectively maintained non-negative pressure upstream from the valve when the RBP was turned on, as shown in the middle panels of Figure 6-1. On the lowest-pressure connections, the upstream supply pressure was negative during most of the time that the ‘system was pressurized.’ This meant that to extract any water from the system, houses required an operating RBP. On these connections, the upstream pressure was not positive, and therefore while the valve’s throttling effect did reduce the magnitude of the negative pressure, it did not restore positive upstream pressure, as seen in the bottom two panels of Figure 6-1. Despite not eliminating all of the negative pressure, the valve reduced much of the suction capacity of the RBPs and therefore caused significant — often greater than 40% — reductions in flow rates. Therefore finding consenting customers with low-pressure connections who would allow a 24-hour study was understandably not possible. The rest of the data presented herein was taken at medium- and high-pressure connections. 117 Figure 6-1: Effect of the valve on different types of house connections 118 6.2 Case study 1: applying the metrics to a house in Old Delhi (J02) To demonstrate the use and meaning of the three performance metrics, a house in Old Delhi (J02) is used, whose pressure history is summarized in Figure 6-2. The top panel shows the pressure history of the house throughout the January field study. Green is the pressure history while the valve was installed; blue is the pressure history without the valve. Each vertical line extending down from the pressure history represents a booster pump cycle, such as those in the middle two panels of Figure 6-1. 6.2.1 Duration of negative pressure To better demonstrate how the duration of negative pressure was calculated, a histogram of the pressure distribution with and without the valve is shown in the middle panel of Figure 6-2; the bin width was 0.5m. Notably, 100% of the negative pressure below -0.75m was eliminated by the valve. The bottom panel shows the difference between the two histograms, control data minus valve data. The valve prevented 93 minutes per day of negative pressure at this connection. Unfortunately, this simple histogram method obscured what happened within the bins. The actual metric used for performance evaluation was the average number of minutes per day at any magnitude of negative pressure. Without the valve, the house averaged 101 minutes per day; with the valve, 63 minutes per day. Despite the visually significant impact of the valve on the pressure history of J02, the valve’s impact was only a moderate 38% reduction in the average daily duration of negative pressure. While this improvement is significant in and of itself, it neglects to account for the magnitude of negative pressure. 6.2.2 CIR The contamination risk, as discussed at length in Section 5.3, is a function of the pressure difference between contaminants outside the pipe and the pressure inside the 119 Pressure Profile 15 Control Valve Pressure (m) 10 5 0 −5 Avg. Min/Day at this Pressure −10 01/22/14 01/23 01/24 01/25 01/26 Date 01/27 01/28 01/29 01/30 Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure 6-2: Negative pressure reduction example from J02 in Old Delhi 120 pipe. The contaminant intrusion ratio (CIR) accounts for this pressure dependency. Contamination also depends on the integrity of the pipe. The CIR accounts for this by normalizing the calculated intrusion rate by the leakage rate, which is also a function of the pipe’s integrity. The CIR therefore calculates the volume ratio of potentially intruded contaminants to water leakage in the service connection pipe. Neither the absolute value of intruded contaminants nor leakage could be calculated from the pressure alone without an assessment of the pipe’s condition. The pipe’s condition is accounted for by an empirical constant, πΏπΆ , assumed (details of this derivation are in Section 5.3.3) to be the same for leakage and intrusion. The connection between the pressure history, the relative volume of leaked water, the relative volume of intruded contaminants, and the CIR, is shown in Figure 63. The top panel again shows the pressure history of J02. The middle two panels show the relative accumulated volume of leakage and intrusion during each half of the trial. These numbers are calculated using Equations 5.12 and 5.13, and have units of π3 . In these panels, it can be seen that while the leakage rates accumulate similarly πΏπΆ between the valve and control, the contaminant intrusion volume accumulates much more quickly without the valve. Finally, the bottom panel graphs the instantaneous CIR during each half of the trial. At the beginning of the control, the pressure was negative and therefore the CIR started at 100%, but this was not shown so as to focus on the more representative CIR range. valve. The total CIR without the valve was 3.2%, but only 0.03% with the This represents contamination reduction of two orders of magnitude. The instantaneous CIR was shown in this section, but elsewhere in this thesis the CIR will always refer to the average CIR, which is to the ratio of daily-averaged intruded contaminants to daily-averaged leaked water. 6.2.3 Health risk Assuming the service connection pipe was exposed to sewage and had a leakage rate of 10%, then using Table 5.7’s correlation of CIR values to health risk levels, house 121 Pressure [m] Pressure History 10 0 −10 01/22/14 01/23 01/24 01/25 01/26 01/27 01/28 01/29 01/30 01/23 01/24 01/25 01/26 01/27 01/28 01/29 01/30 01/23 01/24 01/25 01/26 01/27 01/28 01/29 01/30 01/23 01/24 01/25 01/26 Date 01/27 01/28 01/29 01/30 15000 10000 5000 3 Contaminates Intruded [ LmC ] 3 Water Leaked [ m LC ] Control Valve 20 0 01/22/14 600 400 200 0 01/22/14 CIR 0.04 0.02 0 01/22/14 Figure 6-3: Contamination risk associated with the pressure history of J02 122 J02 had a health risk level of high without the valve and low with the valve. 6.3 Sources of uncertainty in the data While the performance metrics applied to J02 in Old Delhi clearly showed the valve’s impact, the results from other connections showed more variability. Two major sources of error are hypothesized: variability in the baseline pressure history of the house and sensor accuracy. These sources of error are more prominent in the data gathered from intermittently-supplied houses. This section briefly discusses baseline variability and then focuses on sensor accuracy with a case study. 6.3.1 Variability in baseline characteristics of customer connections The crossover design of the field study sought to isolate the impact of the valve on customer connections by comparing the performance metrics with and without the valve. The crossover design relied on the impact of the treatment (the installation of a valve) being larger than the random variation in the control data. For the intermittentlysupplied houses in this sample, the baseline variation was, unfortunately, much larger than the treatment impact. This is discussed more in Section 6.6. 6.3.2 Case study 2: the effects of variability on a house in Lado Sarai (M05) The pressure loggers, used to gather the presented results, measured pressure with a standard deviation of 0.15 meters, as measured in Appendix A. To consider the effects of this variability, a connection from Lado Sarai, M05, is taken as an extreme example of noisy data. The measured data showed that the house experienced more than 10 hours per day of additional negative pressure with the valve installed. Since there is no physical mechanism by which the valve could have made the duration of negative pressure 10 hours longer, the data from this connection serves as an excellent 123 example of what types of uncertainty were in the data and how they could affect the results presented in this chapter. Figure 6-4 shows the pressure profile of M05 over time in the top panel. This house received one hour of high pressure water once per day, spent more than 17.5 hours a day between -0.25 and 0.25 meters of pressure, and had two, two-hour periods per day, where pressure rose to around 0.5 meters. Despite the middle panel of Figure 6-4 showing that the valve made the duration of negative pressure 10 hours longer in the range of -0.35 to -0.05 meters, the bottom panel highlights that the valve made the duration of negative pressure below -1.5 meters shorter by eight minutes per day. This reduction in severe negative pressure, when compared to the baseline negative pressure of approximately four hours per day, results in the low signal-to-noise ratio of eight minutes of signal to four hours of noise. The low signal-to-noise ratio arose because the valve was never intended to prevent all negative pressure, only negative pressure induced by booster pumps. Across nine days of observations at M05, the booster pump’s pressure signature was observed only six times, and only once trying to draw water outside of the high-pressure time period. Instead of just a low signal-to-noise ratio however, the data suggests that the presence of the valve made the duration of negative pressure longer. Overlaying the pressure history of each day and considering in detail the observed pressures around zero, Figure 6-5 shows a mean shift between control and treatment. Since it is unlikely that valve would have caused an increase in the duration of negative pressure, two alternative causes are hypothesized: drift in the pressure logger, which would have been within 2 standard deviations (±0.3m) of the observed variance in the pressure loggers detailed in Appendix A; or, the system pressure during the second five days of observation was higher than during the first five days. Since the pressure uniformly shifted even during times when the system was turned off, it is likely that this difference was caused by drift in the pressure logger. Whatever the reason for the effect, connections such as M05 were very sensitive to minor changes in either the actual or the measured pressure while the system was 124 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 Avg. Mins/Day at this Pressure Avg. Mins/Day at this Pressure −5 03/30/14 03/31 04/01 04/02 04/03 04/04 04/05 04/06 04/07 Date Pressure Histogram Detailed at Zero Pressure 04/08 04/09 Control Valve 400 300 200 100 0 −0.5 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 Pressure Bin (m) Severe Negative Pressure Histogram 0.3 0.4 0.5 Control Valve 1.2 1 0.8 0.6 0.4 0.2 −4.5 −4 −3.5 −3 −2.5 −2 Pressure Bin (m) −1.5 −1 −0.5 0 Figure 6-4: How the valve caused a longer duration of negative pressure at M05 125 Control Pressure Profile Control Valve 1.5 Pressure (m) 1 0.5 0 −0.5 −1 01/00/00−00:00 03:00 06:00 09:00 12:00 Date 15:00 18:00 21:00 Figure 6-5: Evidence of a mean shift in the sensor at M05 turned off. This sensitivity is not a mathematical quirk, but arises out of the physics of the system; systems that spend many hours per day unpressurized can easily become contaminated if the contaminants are slightly pressurized, or if the system becomes negatively pressurized. This is a severe risk associated with intermittently-operated systems that the valve was never intended to correct; the valve can only reduce the negative pressure caused by RBPs. The root cause of this sensitivity is therefore the large portion of each day during which the pressure is very close to zero. To account for this sensitivity in the rest of the results presented in this chapter, a Monte Carlo simulation was conducted with N=1000. A normally distributed, zeromean, mean shift was applied to each half of the crossover data, with a standard deviation of π = 0.15m. I.e. for a specific house, the control and the treatment pressure histories each received a different random error. More details about the simulation method are included in Section 5.4.1. The results presented in the rest of this chapter are the mean values of this Monte Carlo simulation; error bars spanning the 1st and 99th percentile of the simulation are shown in red and error bars spanning the 5th and 95th percentile are shown in blue. 126 6.4 Total aggregate impact of the valve The valve’s function was to reduce (or eliminate) the magnitude and duration of upstream negative pressure. Seventeen of the 19 observed houses averaged more than one minute per day of pressure less than zero. As was in the case study of M05, several of the intermittently-connected houses had long durations of pressure close to atmospheric pressure, which were very sensitive to small shifts in the data. To highlight the smaller, but important changes in severe negative pressure, this section presents the duration of negative pressure below several thresholds. The valve most consistently eliminated negative pressure below -1 meter. 6.4.1 Duration of negative pressure In the top half of Figure 6-6, each data point represents one house; its x coordinate is the length of time that the house averaged below -2m without the valve, and the y coordinate is the length of time that the house spent below -2m with the valve installed. The neighborhood location and connection characteristics of each data point are tabulated by label in Table 5.2. For confidentiality reasons, the addresses of these houses were not published, but more information is available upon request of the author. The valve eliminated 100% of observed upstream negative pressure less than -2m, which occurred for an average (across all 19 houses) of 38 minutes per house, per day. Similarly, in the bottom half of Figure 6-6, the valve prevented 96% of all negative pressure less than -1m, which occurred for an average of 52 minutes per house, per day. The valve’s reduction of pressure between -1m and 0m was less clear, and the impact on all negative pressure (below 0m), was statistically inconclusive. As shown in Figure 6-7, eight of the 19 data points suggest we must accept the null hypothesis (blue dashed line) that the valve did not reduce negative pressure. Data points with magenta and black error bars are continuously-supplied houses, while the standard blue and red error bars are intermittently-supplied houses. 127 Because of the stark Duration with Valve (Avg. Min/Day) Sensitivity of Negative Pressure Duration <−2m 80 60 40 20 Duration with Valve (Avg. Min/Day) 0 J08 M04 M05J05 J03 J01 J07 J06 J02 M03 0 50 M01 100 150 200 Duration without Valve (Avg. Min/Day) Sensitivity of Negative Pressure Duration <−1m 250 100 50 M01 M07 J08 J06 M03 J09 J03J01 J05 J02 J07 0 0 50 100 150 200 250 300 350 Duration without Valve (Avg. Min/Day) Figure 6-6: Severe negative pressure reduction: valve consistently reduced negative pressure less than -2m and -1m 128 contrast between the continuously and intermittently supplied connections, they are considered independently in Sections 6.5 and 6.6. While it must be concluded that the overall impact of the valve on the total duration of all negative pressure was inconclusive, the valve did reduce 96% of negative pressure less than -1 meter — an average of 52 minutes per house per day. 6.4.2 CIR Figure 6-8 shows the total impact the valve had on the CIR; the x-axis was used as the control direction (the CIR without the valve), the y-axis was used as the treatment direction (the CIR with the valve). While the uncertainty was still very large, 13 of 19 data points had mean values that favored rejecting the null hypothesis. This suggests that the valve did likely improve contamination, but the actual impact was vividly different between continuously and intermittently-supplied connections; they are therefore treated separately in Sections 6.5 and 6.6. 6.5 Aggregate impact of the valve on continuouslysupplied houses As seen in the previous section, the performance of the valve differed significantly from continuously to intermittently-supplied houses. The mean impact of the valve on the 12 continuously-supplied houses was: a reduction in the duration of negative pressure by 82 minutes per house, per day; a 74% reduction in CIR values; and, the shifting of seven customers out of the the low and safe high and very high health risk categories into categories. This section elaborates on the valve’s effect on each of these three metrics. The effect of the valve on continuously-supplied houses was clearer than intermittentlysupplied houses for two related reasons: first, the control measurements of negative pressure more effectively identify the booster pump problem, as without booster pumps, the supply pressure would never be negative; and, second, were the valve 129 Null Hypothesis 1200 M03 J08 1000 M06 M05 Duration with Valve (Avg. Min/Day) M04 M07 800 600 400 M02 200 M01 J01 0 0 Figure 6-7: J05 J02 J06 J03 200 J07 400 600 800 1000 Duration without Valve (Avg. Min/Day) 1200 Inconclusive impact of duration of negative pressure; clear differences between intermittent connections (magenta and black) and continuous connections (blue and red); null hypothesis that the valve has no impact is the blue dashed line 130 3.5 Null Hypothesis 3 M03 CIR with Valve 2.5 2 M06 1.5 1 J08 M05 M04 0.5 M02 0 M07 M01 0 Figure 6-8: 0.5 1 1.5 2 2.5 CIR without Valve 3 3.5 4 4.5 Inconclusive impact on CIR; intermittent connections (magenta and black) are dwarfed by the continuous connections (blue and red); null hypothesis that the valve has no impact is blue dashed line 131 450 Null Hypothesis 400 Duration with Valve (Avg. Min/Day) 350 300 250 200 M01 150 100 J02 J05 J06 J07 J03 J01 0 J09 0 100 200 300 400 Duration without Valve (Avg. Min/Day) 50 Figure 6-9: 500 Time spent at negative pressure on continuous connections; red error bars span the 1st and 99th percentile of results from the Monte Carlo simulation; blue error bars, the 5th and 95th to work perfectly on these houses, the total intrusion risk could be brought to zero. 6.5.1 Reduction in the duration of negative pressure The valve’s impact on the duration of negative pressure is shown in Figure 6-9. Nine of twelve data points have mean values which favor rejecting the null hypothesis. All three that support the null hypothesis are very close to zero; the valve increased the duration of negative pressure for only a total of eight minutes between these three houses. The mean values therefore clearly indicate that overall the valve significantly reduced the duration of negative pressure. To better understand the certainty of these results, two observations should be made about the span of the error bars in Figure 6-9: first, the range in the x or control direction is smaller than the y or treatment direction; and second, the range in the y or treatment direction frequently spans up to the null hypothesis line. A narrow 132 range of error bars in the control directions indicates that most of the negative pressure experienced at these connections was below -0.45 meters. Adding a random noise with three standard deviations equal to 0.45m still did not shift most of these observed negative pressures into the positive region. This allows the confident conclusion that negative pressure was a problem at continuously-supplied houses. If the valve had functioned perfectly, it would have throttled all negative pressure to exactly zero. Adding noise to this expected pressure signal would result in half of the data points being shifted back to negative pressure values, leaving the duration of negative pressure unchanged. Therefore, it was expected that the y-error bars in Figure 6-9 would stretch up from the x-axis to the null hypothesis line. Where this was not the case, the data may have been skewed (in either direction) by a negative pressure event not caused by booster pumps or the valve’s set point may have drifted up to a pressure higher than +0.45m. Both these possibilities were seen in the data and are discussed below. The y-error bar of J07 is significantly shorter than expected. Figure 6-10 shows J07’s pressure history in detail; during the first two days of the control measurements, there was a significant drop in system pressure not related to booster pumps, and therefore this data should be excluded. This type of event in the control data inflates the perceived effect of the valve. Beginning the control data on January 22 instead of January 20, 2014, resulted in Figure 6-11, which was closer to expectations. For completeness, the pressure histories of all 19 connections are included in Appendix C. Another non-booster pump event skewed the results of M01, whose pressure history is shown in Figure 6-12. Whereas it was typical for the system pressure to drop to almost zero at midday, on April 3rd (04/03), it dropped to -1m. This non-booster pump event created negative pressure that the valve could not prevent and deflated its perceived effectiveness. This explains in part why the data point was elevated and the y-bar did not reach down to the x-axis. However, another explanation is needed for why the y-error bar did not reach the null hypothesis line. On several occasions at M01, the valve throttled the booster pump to a set point of between positive 0.5 and 1m. One such example is shown in Figure 6-13. While 133 Pressure Profile 20 Control Valve 15 Pressure (m) 10 5 0 −5 −10 01/20/14 Figure 6-10: 01/21 01/22 01/23 01/24 Date 01/25 01/26 01/27 01/28 Detail of pressure history at J07; highlights the anomaly during the control measurement Null Hypothesis 300 Duration with Valve (Avg. Min/Day) 250 200 150 100 M01 J02 50 J05 J03 J06 J09 J01 J07 0 0 100 200 300 400 Duration without Valve (Avg. Min/Day) 500 600 Figure 6-11: Time spent at negative pressure on continuous connections; adjusted for J07 134 Pressure Profile 10 Control Valve 8 Pressure (m) 6 4 2 0 −2 −4 03/31/14 04/01 04/02 04/03 04/04 04/05 04/06 04/07 04/08 04/09 Date Figure 6-12: Detail of pressure history at M01; highlights the anomaly during the valve measurement on midday 04/03 this higher set point made the reduced duration of negative pressure more robust to the magnitude of error used in the Monte Carlo simulation, it also caused higher reductions in flow rate. This higher set-point was observed in several of the alpha prototypes as was discussed in Section 4.3.6. Summarizing the observed effect of the valve on continuously-supplied houses, Table 6.1 quantifies the mean, median, maximum, and minimum impact of the valve on the duration of negative pressure. From the mean values, it can be confidently concluded that the null hypothesis should be rejected and that the valve does indeed reduce the duration of negative pressure. 6.5.2 Reduction in CIR Figure 6-14 compares the CIR without the valve to the CIR with the valve at each continuously-supplied connection and demonstrates that the valve caused a very significant reduction in the contamination risk across every continuously-supplied house. Even at connections where the valve did not prevent all the time spent at negative pressures, such as J02, J05 and J06, the valve forced the negative pressures close enough to zero to significantly limit their associated contamination risk. The range 135 Pressure Profile 4 Control Valve 3.5 3 Pressure (m) 2.5 2 1.5 1 0.5 0 −0.5 −1 04/01/14−14:00 15:00 16:00 17:00 Date 18:00 19:00 20:00 Figure 6-13: Extreme detail of pressure history at M01; highlights the valve throttled to a higher pressure than expected during booster pump events Table 6.1: Summary of negative pressure reduction on continuously supplied houses; -∞ is treated as -100% for mean calculation Duration of Negative Mean Median Maximum Minimum (mins/day) (mins/day) (mins/day) (mins/day) 109 64 481 0 Valve 27 13 145 0.1 Single house difference 82 44 351 -3.2 37% 70.5% 99.4% -∞ Pressure; N=12 Control (No valve) (πΆπ − ππ ) Single house percent π difference (1 − π ) πΆπ 136 CIR with Valve M01 0.04 0.02 0 J06J05J02 0 0.05 Null Hypothesis J07 0.1 0.15 0.2 CIR without Valve 0.25 0.3 Figure 6-14: Valve’s ability to change CIR values for continuously pressurized houses; CIR is extremely robust to noise in this data set Table 6.2: Summary of contamination reduction on continuously supplied houses; a negative value means the CIR increased when the valve was installed CIR Values; N=12 Control (No valve) Valve Single house difference (πΆπΌπ control (π) Median Maximum Minimum 0.04 0.02 0.2 0 0.004 1 × 10−5 0.05 0 0.03 0.02 0.2 −5 × 10−6 74% 97% 100% -68% − πΆπΌπ valve (π)) Percentage Difference (1 Mean − πΆπΌπ valve (π)/πΆπΌπ control (π)) of calculated CIR values for continuously supplied houses is summarized in Table 6.2. Notably, the mean effect was a 74% reduction in contamination risk and the maximum observed effect was a CIR reduction of difference of 0.2 at M01. The error bars in Figure 6-14 demonstrate that the valve’s reduction in contamination was robust to noise. More generally, the CIR metric was always less sensitive to noise than the duration of negative pressure metric because it took into account the magnitude of negative pressure. For example, a small shift in the pressure signal from +0.01m to -0.01m could significantly skew the duration of negative pressure metric, but would have little effect on the CIR metric. 137 Table 6.3: Reduction in health risk at continuously supplied houses Health Risk Maximum CIR with 10% Houses in Houses leakage & sewage exposure without Valve with Valve <0.01% 3 8 0.01-0.1% 2 3 Intermediate 0.1-1% 0 0 High 1-10% 5 1 Very high >10% 2 0 Level Safe: meets requirements Low 6.5.3 Reduction in health risk The worst-case assumptions and method detailed in Section 5.3.5 were used to determine the health risks associated with the CIR values found; the result was Table 6.3. Six of seven connections with safe very high and high risk levels improved to levels. The one connection that remained in the high low and risk category even with the valve was M01, whose non-booster pump negative pressure was already discussed in Section 6.5.1. During the control, M01 had a CIR of 25% and with the valve it had a CIR of 5%. Therefore, even at M01, the valve still reduced contamination by almost an order of magnitude and the health risk by a risk category. The valve’s clear and significant impact on negative pressure, contamination, and health risks at continuously supplied houses urgently impels the further development of this valve. 6.6 Aggregate impact of the valve on intermittently supplied houses As discussed in Section 2.1.3, one of the major risks associated with the intermittent supply of water is contaminant intrusion during un-pressurized times. When the system is first turned on, the contaminants which have been percolating into the pipe while the system was turned off are flushed out; during this flushing, the water has measurably lower quality [4]. The tested valve cannot prevent this contamination 138 source; however, the metrics used to evaluate the valve’s performance do not differentiate between negative pressure during times when the system is turned off, and booster-pump-induced negative pressure. Because of this, the intermittently-supplied connections in this study had very low signal-to-noise ratios, which prevented any conclusion from being made about the valve’s overall impact on intermittently-supplied houses. This signal-to-noise ratio is discussed more in Section 6.6.1. To improve the signal-to-noise ratio, the pressure and contamination at these houses are considered only when the system is turned on (during supply hours). Even after isolating for supply hours, the signal-to-noise ratio was still low because only three of seven houses had booster-pump-induced negative pressure lasting more than ten minutes per day. Accordingly, the valve’s reduction in the duration of negative pressure during supply hours was statistically inconclusive, despite having an average reduction of 22%. of contaminant intrusion. The valve did, however, conclusively reduce the risk With the valve, the CIR was reduced by an average of 26%; unfortunately this improvement only correlated to an improvement in the risk threshold of one household, which went from high to intermediate risk. 6.6.1 Intermittent connections require data filtering Filtering the data after the trial was required to improve the signal-to-noise ratio. The three reasons the intermittently-supplied houses in this trial had low signal-to-noise ratios were: 1. particular to this trial, four of seven of the intermittently-supplied customers used their booster pumps for less than 10 minutes per day; the total booster pump usage is summarized in Table 6.4; 2. particular to this trial, the duration of the control and treatment observation windows was not long enough to account for the baseline variability in negative pressure and contamination risk; and, 3. all intermittently-supply customers had stretches of time where negative pressure and contamination risk were not from booster pumps, but from the system’s 139 Table 6.4: Booster pump induced pressure characteristics on intermittent houses House Duration of Negative Pressure RBP’s Suction Supply (minutes / day) Pressure (m) Pressure (m) ≈ 20 J08 ≈ -2.5 ≈ 3 M02 None N/A N/A M03 ≈ 100 ≈ 25 ≈5 ≈2 ≈3 ≈ ≈ ≈ ≈ ≈ ≈1 ≈0 ≈ 10 ≈7 ≈ 12 M04 M05 M06 M07 -3 -3 -4 -4 -4 depressurization. The design of the crossover field study was based on the assumption that a given house would have a similar pressure history from one day to another and therefore the treatment effect could be found by considering the difference between the control and treatment measurements. To verify this assumption, Figure 6-15 plots the duration of negative pressure during the first half of the control measurements for intermittentlysupplied houses against the second half. The figure demonstrates that the baseline variation was extremely high. To improve upon this, future field work must strive to lengthen the treatment and control observation windows to average out this source of noise. To visualize the signal-to-noise ratio without isolating for supply hours, Figure 616 plots the expected result from a perfect valve as a red dot on top of the duration of negative pressure as measured with and without the valve. The valve’s ability to reduce the duration of negative pressure could not be verified. Even if the valve had performed perfectly, its impact would not have been significant compared to other sources of negative pressure. Using the CIR metric instead resulted in the similarly inconclusive Figure 6-17. Together, these figures force the conclusion that the null hypothesis must be accepted; the valve did not significantly affect contamination or the duration of negative pressure on intermittently-supplied houses. The rest of this section focuses on the valve’s ability to reduce negative pressure and contamination 140 Duration during the second half of the control (Avg. Min/Day) 1200 M03 M06 1000 M07 800 M05 J08 600 M04 400 M02 200 0 −200 0 200 400 600 800 1000 1200 Duration during the first half of the control (Avg. Min/Day) 1400 Figure 6-15: The duration of negative pressure in the first half of the control observation compared to the second half shows that the variability in the supply pressure to intermittent houses was larger than the effect of the valve; consistent data would sit on the null hypothesis line during the time when the system is turned on — during supply hours. 6.6.2 Duration of negative pressure during supply hours By isolating the valve’s performance during supply hours, using the method described in Section 5.4.3, the signal-to-noise ratio was slightly improved. Figure 6-18 shows the valve’s impact on the duration of negative pressure; the theoretically expected performance of a perfect valve (see Table 6.4) was superimposed as red dots. While four of seven mean values suggest the null hypothesis must be accepted, only one of the three connections (M04) which had more than ten minutes per day of negative pressure suggests accepting the null hypothesis. Both J08 and M03 have mean values that suggest that the null hypothesis must be rejected; no part of J08’s uncertainty bars span the null hypothesis line. It is therefore likely that while the valve did not have any overall impact on the duration of negative pressure during supply hours at intermittently-connected houses, the valve did improve the negative pressure duration on connections that had at least ten minutes per day of booster-pump-induced 141 1200 M03 J08 Duration with Valve (Avg. Min/Day) 1000 M05 M06 M04 800 M07 600 Null Hypothesis Theoretically Expected Results 400 M02 200 0 0 200 400 600 800 1000 Duration without Valve (Avg. Min/Day) 1200 Figure 6-16: Valve’s impact on negative pressure reduction on intermittently-supplied houses is inconclusive because of very low signal-to-noise ratios; a perfect valve’s expected result is overlaid in red 142 5 Null Hypothesis Theoretically Expected Performance 4.5 4 CIR with Valve (Avg. Min/Day) 3.5 3 2.5 M03 2 1.5 M06 J08 1 M04 M05 M07 0.5 M02 0 0 0.5 1 1.5 2 2.5 3 3.5 CIR without Valve (Avg. Min/Day) 4 4.5 5 Figure 6-17: Valve’s impact on CIR on intermittently-supplied houses is inconclusive because of very low signal-to-noise ratios; a perfect valve’s expected result is overlaid in red 143 negative pressure. Figure 6-19a highlights why the valve was not able to reduce all of the time spent at negative pressure at M04: after the characteristic sudden drop and then sudden rise in pressure associated with a booster pump cycle, the pressure at the customer connection continued to be negative. M04 used booster pump Configuration 2 where the booster pump is placed upstream from the sump. If the outlet into the sump had been below the elevation of the distribution main, after the booster pump had turned off, the water would have continued to be pulled into the sump. If this was the case, the beta prototype used at that connection was either not sensitive enough, or did not seal tightly enough to prevent this small magnitude (-0.5m to 0m) negative pressure. Figure 6-19b overlays the evening pressure recordings from M04. While the valve did clearly prevent the severe negative pressure caused by the booster pump action, variability in the starting time and duration of the positive pressure cycles are also visible. For several days when the valve was installed, there was a delayed start to supply hours, which was the most likely cause for the measured increase in negative pressure duration. This variation in the baseline behavior of the system also affected the CIR values. The quantitative performance of the valve was equally inconclusive, as summarized in Table 6.5. The only positive metric was that the valve did reduce the negative pressure duration by an average of 22%. 6.6.3 CIR during supply hours The valve reduced the contamination risk during supply hours at intermittentlysupplied connections. Figure 6-20 shows the valve’s impact on the CIR; the expected results from a perfect valve are superimposed as red dots. Six out of seven mean CIR values support rejecting the null hypothesis; therefore, the valve did reduce the risk of contamination during supply hours. As in the case of the duration of negative pressure, J08 and M03 showed significant benefits from the valve. Table 6.6 quantifies that the presence of the valve did improve the CIR. The results 144 Null Hypothesis Perfect Valve Performance 200 M03 M04 Duration with Valve (Avg. Min/Day) 150 100 M02 M06 50 M07 M05 J08 0 50 100 150 Duration without Valve (Avg. Min/Day) 200 Figure 6-18: Negative pressure reduction during supply hours on intermittent houses; red dots indicate theoretically ideal performance; red bars are the 99% CI, blue the 95% CI 145 2 Pressure (m) 1 0 −1 −2 −3 −4 04/03/14−16:00 16:30 17:00 17:30 18:00 18:30 Date (a) Detail of negative pressure at M04; booster pump is followed by small magnitude negative pressure of unknown origin Pressure History Overlaid M04 1.5 1 0.5 Pressure (m) 0 −0.5 −1 −1.5 −2 −2.5 −3 −3.5 Control Valve 01/00/00−16:00 16:30 17:00 17:30 Time 18:00 18:30 19:00 (b) For several days when the valve was installed, there was a delayed start to supply hours, which increased the duration of the non-booster-pump negative pressure and skewed the results Figure 6-19: Investigating the unexpected results form M04 146 0.9 Null Hypothesis Theoretically Expected Performance 0.8 0.7 CIR with Valve (Avg. Min/Day) 0.6 M03 0.5 0.4 0.3 M04 0.2 0.1 M05 M07 M06 M02 0 J08 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 CIR without Valve (Avg. Min/Day) 0.8 0.9 1 (a) All seven connections Null Hypothesis Theoretically Expected Performance 0.3 CIR with Valve (Avg. Min/Day) 0.25 0.2 0.15 0.1 M07 M05 0.05 M06 M02 J08 0 0 0.05 0.1 0.15 0.2 0.25 CIR without Valve (Avg. Min/Day) 0.3 (b) Detail of lower five CIR values Figure 6-20: CIR reduction during supply hours on intermittent houses; red dots are expected values, red error bars are 99% range and blue error bars are the 95% range 147 Table 6.5: Duration of negative pressure at intermittently-supplied houses during supply hours; a negative difference implies the valve made negative pressure worse Duration of Negative Pressure; N=7 Mean Median Maximum Minimum (mins/day) (mins/day) (mins/day) (mins/day) Control (No valve) 81 50 212 31 Valve 77 48 174 10 4 -5 39 -38 22% 22% 80% -47% Single house difference (πΆπ − ππ ) Single house percent π difference (1 − π ) πΆπ Table 6.6: Summary of contamination reduction on intermittently supplied houses, during supply hours; the negative difference in CIR means the presence of the valve increased the CIR value N=7 Mean Median Maximum Minimum Control (No valve) 0.19 0.08 0.78 0.02 Valve 0.13 0.07 0.50 0.01 Single house difference 0.05 0.01 0.28 −2 × 10−3 26% 12% 82% -3% (πΆπΌπ control (π) − πΆπΌπ valve (π)) Percentage Difference (1 − πΆπΌπ valve (π)/πΆπΌπ control (π)) are not as large as those from continuously-supplied houses, but they are nevertheless significant. The mean improvement in CIR because of the valve was 0.05, representing a 26% improvement in the CIR. In the best example of the valve’s benefit, at M03, the absolute CIR was lowered by 0.28. 6.6.4 Health risk during supply hours While the presence of the valve did improve the CIR, in order to correspond to decreases in the health risks, the CIR must decrease by an order of magnitude. Table 6.7 shows that only one of the seven houses had an improvement in their health risk category. While a project which would decrease the health risks of one in seven houses might still be worth pursuing, it is the author’s strong hypothesis that the incon- 148 Table 6.7: Reduction in health risk at intermittently supplied houses, during supply hours Health Risk Maximum CIR with 10% leakage Houses in Houses & sewage exposure Control with Valve <0.01% 0 0 0.01-0.1% 0 0 Intermediate 0.1-1% 0 1 High 1-10% 5 4 Very high >10% 2 2 Level Safe: meets requirements Low clusive data in this section is misleadingly underwhelming. Because of the sampling bias that houses needed to give informed consent for an experiment on their house, and because houses could opt out at anytime, the intermittently supplied houses that frequently used their booster pumps did not enter or stay in the trial long enough to qualify for inclusion. Had such houses been part of the study, the fraction of negative pressure problems caused by booster pumps would have been larger and, therefore, the valve’s impact would have been more significant. The top priority for future field work is to better characterize the pressure histories of representative intermittently-supplied connections. If the connections summarized in this section are representative of the typical intermittently-supplied connections, the valve’s utility to such connections is dubious because a perfect valve would only be able to solve a small fraction of the pressure and contamination problem. Conversely, if the sampled intermittently-connected houses are not representative (as the author has hypothesized), then the valve would still be able to address one significant source of the contamination problem and would therefore be worth pursuing. 149 6.7 Valve’s aggregate impact whenever water was being supplied This section combines the results from Sections 6.6 and 6.5 to present the valve’s aggregate impact on all tested connections whenever water was being supplied; continuouslysupplied houses were examined in their entirety, while observations from intermittentlysupplied houses were cropped using the method described in Section 5.4.3. 6.7.1 Duration of negative pressure The combined effect on the duration of negative pressure is shown in Figure 6-21. The error bars span the 1st and 99th percentile, black represents continuously-supplied connections, while magenta represents intermittently-supplied ones. Despite the inconclusive subset of intermittently-supplied houses, when taken as a whole, the data suggests rejecting the null hypothesis. Table 6.8 quantifies the valve’s average im- pact on negative pressure; notably, the presence of the valve prevented an average of 53 minutes per day, per connection of negative pressure, with a maximum observed benefit of almost six hours per day of negative pressure prevented at J07 and M01. While the valve significantly improved negative pressure, it must be noted that the valve’s impact on intermittently-supplied connections was much less conclusive and requires immediate further study. 6.7.2 CIR Figure 6-22 shows the valve’s combined impact on the CIR across all connections while water was being supplied. While large uncertainty was still associated with the intermittently-supplied connections, the overall results were clear: the valve significantly reduced the contamination risk. For a more detailed view of the continuously supplied connections in Figure 6-22, the reader is directed to Figure 6-14. Table 6.9 quantifies the significant benefits of the valve: the contamination risk was reduced by a mean of 43% and a median of 80%. 150 Null Hypothesis 350 Duration with Valve (Avg. Min/Day) 300 250 200 M03 M04 150 M01 100 M02 M06M07 J02 M05 J06 J05 J08 J03 J09 J01 0 0 100 50 J07 200 300 400 Duration without Valve (Avg. Min/Day) 500 Figure 6-21: During supply times, the valve decreased the duration of negative pressure, but the intermittent connections are an inconclusive subset of this data; in pink are intermittently-supplied connections, in black are continuously-supplied connections; the error bars span the 1st and 99th percentile of the Monte Carlo simulation Table 6.8: Summary of negative pressure reduction during supply hours; a negative difference means an increase in the pressure duration with the valve; −∞ is taken as -100% for the mean calculation Negative Pressure Mean Median Maximum Minimum Duration (N=19) (mins/day) (mins/day) (mins/day) (mins/day) Control (No valve) 99 57 484 0 Valve 45 35 175 0.1 Single house difference 53 22 351 -38 23% 49% 99% −∞ (πΆπ − ππ ) Single house percent π difference (1 − π ) πΆπ 151 1 Null Hypothesis 0.9 0.8 CIR with Valve (Avg. Min/Day) 0.7 0.6 M03 0.5 0.4 0.3 M04 0.2 M05 M06 M07 0.1 0 Figure 6-22: M02 J02 0 M01 J07 0.1 0.2 0.3 0.4 0.5 0.6 0.7 CIR without Valve (Avg. Min/Day) 0.8 0.9 During supply times, the valve reduced the CIR, but the intermit- tent connections are a less conclusive subset of this data; in pink are intermittentlysupplied connections; in black are continuously-supplied connections; the error bars span the 1st and 99th percentile of the Monte Carlo simulation; for a closer look at the continuously supplied data points, see Figure 6-14 152 Table 6.9: Summary of contamination reduction during supply hours; a negative difference implies an increase in CIR with the valve; −∞ is taken as -100% for the mean calculation CIR Values (N=19) Mean Median Maximum Minimum Control (No valve) 0.10 0.02 0.78 0 Valve 0.06 5 × 10−4 0.50 0 Single house difference 0.04 0.02 0.28 −3 × 10−3 43% 80% 100% −∞ (πΆπΌπ control (π) − πΆπΌπ valve (π)) Percentage Difference (1 − πΆπΌπ valve (π)/πΆπΌπ control (π)) Table 6.10: Reduction in health risk during supply hours Health Risk Maximum CIR with 10% leakage Houses in Houses & sewage exposure Control with Valve <0.01% 3 9 0.01-0.1% 2 2 Level Safe: meets requirements Low Intermediate 0.1-1% 0 1 High 1-10% 10 5 Very high >10% 4 2 6.7.3 Health risk To reduce the health risks associated with contamination, contamination has to be reduced by an order of magnitude. Table 6.10 documents the number of houses in each risk threshold, calculated according to the method in Section 5.3.5. Notably, the valve removed seven of fourteen houses from the rendered six of those seven in the safe very high and high risk category and category. Therefore, even amidst the baseline uncertainty and the infrequent booster-pump usage in the intermittently-supplied connections, the reduction in CIR and health risks clearly demonstrates the valve’s utility and importance. 153 Table 6.11: Given an assumed booster pump suction capacity, this table summarizes the minimum required supply pressure so that when the valve is installed, customer flow rates are not reduced more than the amount in the column heading Typical Booster Pump 10% Flow 25% Flow 40% Flow Suction Capacity Reduction Reduction Reduction -0.5m 2.1m 0.6m 0.3m -1m 4.3m 1.3m 0.6m -2m 8.5m 2.6m 1.1m -3m 12.8m 3.9m 1.7m -4m 17.1m 5.1m 2.3m -5m 21.3m 6.4m 2.8m 6.8 The necessity of a pilot study and recommendations for its scoping The crossover study of individual house connections presented in this results section, was not able to verify the valve’s aggregate impact on water and pressure distribution equality. To investigate this impact, a pilot study, in which one valve would be installed at each connection in a small hydraulically-isolated area, is required. Executing such a pilot study, or even a longer crossover study would only be feasible where consumer responses to the valve’s installation could be managed. Whenever negative pressure is prevented, flow rates are reduced, and customers are sometimes unhappy. If a water utility were able to determine what reduction in flow rate would be acceptable to their customers, then given an assumption about the strength of customers’ booster pumps, Table 6.11 summarizes the minimum supply pressure that must be provided by the utility to ensure that flow rates are not limited more than the acceptable amount. If for example, booster pumps are presumed to induce -2m of negative pressure and customers will accept a 25% flow reduction, then the valve should be installed only in areas with a minimum supply pressure of +2.6m. The flow rate reductions could be further mitigated if the valves were installed in a isolated neighborhood, where the system pressure could be independently controlled and temporarily increased to compensate for the flow reduction caused by the valves. 154 In such a pilot study, the valve’s benefits would be most evident where the supply pressure is below +6m of pressure for at least half of the supply time, thereby allowing booster pumps to induce negative pressure. An ideal pilot study neighborhood would therefore be one which: 1. is hydraulically isolated, 2. has good relationships between the residents and the water utility, 3. has a supply pressure below +6m for at least 50% of the supply hours, 4. has independently controllable pressure, 5. always has a minimum pressure of +2.5m during supply hours, and 6. is as small as possible. 155 THIS PAGE INTENTIONALLY LEFT BLANK 156 Chapter 7 Conclusions and future work This thesis sought to improve water quality in intermittently-supplied water networks by preventing residential booster pumps from inducing negative pressure in the water supply pipe, which can cause contaminant intrusion. To effectively do so, a patentpending, self-actuating, full-bore, back-pressure regulating valve was designed and tested in New Delhi, India. A crossover study was conducted to isolate and verify the valve’s ability to reduce the magnitude and duration of negative pressure, as well as decrease the associated water contamination risks at intermittently- and continuouslysupplied connections. The valve proved most effective at continuously-supplied connections, preventing an average of 82 minutes per day, per house of negative pressure. The valve also reduced the contamination risk at each connection by a median of 97%, enough to cause significant improvements in the contamination-associated health risks. The results from intermittently-supplied connections were much less clear because: four of the seven intermittently-supplied connections in the study did not have instances of booster-pump-induced negative pressure longer than 10 minutes per day; the booster-pump-induced risk at the remaining three connections was still only a small fraction of the total risk; and high variability in the baseline pressure history of these connections obscured the effects of the valve. Therefore these seven connections had a low signal-to-noise ratio, and most conclusions needed to be qualified by ‘while the system was pressurized.’ 157 Combining the results from continuously- and intermittently-supplied connections, this study showed that: β the valve was most effective at limiting severe negative pressure, preventing 96% of all pressure less than -1 meter; β the valve reduced the duration of negative pressure by an average of 53 minutes per day, per connection while the system was pressurized; β the valve reduced the risk of contamination by a median of 80% while the system was pressurized; β the negative pressure caused by booster pumps posed a significant health risk while the system was pressurized; and β the valve not only took seven of the fourteen houses out of the very high high safe health risk levels, but also brought six of those seven into the and category. During the process of scoping, designing, and testing the valve, several opportunities to improve the design, to improve the test method, and for alternate projects were encountered; they are summarized briefly in Appendix B. Three key future steps are: 1. The valve was designed to prevent negative pressure by throttling to exactly atmospheric pressure; this left no room for variability, drift, or error in this set point. Future iterations of the valve must compensate for this variability by throttling to a higher set point. 2. The seven tested, intermittently-supplied connections had the majority of their contamination risk arise from non-booster-pump events, even while the system was pressurized. The author has hypothesized that this resulted from a sampling bias, and that a representative sample would show booster-pump events playing a larger role in contamination. This hypothesis must be verified immediately. 158 3. The aggregate benefits of the valve must be verified in a pilot study where one valve would be connected to each house in a hydraulically-isolated neighborhood. Such a study would gather the requisite proof that the valve not only reduces contamination but also increases supply pressure and distribution equity, justifying its application on a broader scale. 159 THIS PAGE INTENTIONALLY LEFT BLANK 160 Bibliography [1] World Health Organization and UNICEF, “Data resources and estimates of the WHO/UNICEF joint monitoring programme for water supply and sanitation,” 2010. [Online]. Available: http://www.wssinfo.org/data-estimates/table/ [2] D. McKenzie and I. Ray, “Urban water supply in india: status, reform options and possible lessons,” Water Policy, vol. 11, no. 4, pp. 442–460, 2009. [Online]. Available: http://dx.doi.org/10.2166/wp.2009.056 [3] Asian Development Bank and India. Ministry of Urban Development, Benchmarking and data book of water utilities in India. Manila: 2007 Asian De- velopment Bank, 2007. [4] E. an vol. Kumpel and intermittent 47, no. 14, K. L. and Nelson, “Comparing continuous pp. microbial water quality in piped water supply,” Water Research, Sep. 2013. [Online]. Available: 5176–5188, http: //dx.doi.org/10.1016/j.watres.2013.05.058 [5] P. S. Kelkar, S. P. Andey, S. K. Pathak, and K. G. Nimbalkar, “Evaluation of water distribution system for water consumption, flow pattern and pressure survey during intermittant vis-à-vis continuous water supply in panaji city,” Indian Water Works Association, Journal of vol. 34, pp. 27–36, 2002. [6] K. Vairavamoorthy and K. Elango, “Guidelines for the design and control of intermittent water distribution systems,” Waterlines, vol. 21, no. 1, pp. 19–21, Jul. 2002. [Online]. Available: http://dx.doi.org/10.3362/0262-8104.2002.041 [7] E. Lee and K. Schwab, “Deficiencies in drinking water distribution systems in developing countries,” J water health, vol. 3, pp. 109–127, 2005. [Online]. Available: http://www.iwaponline.com/jwh/003/jwh0030109.htm [8] T. R. Lindley and S. G. Buchberger, “ Assessing intrusion susceptibility in distribution systems,” Journal of American Water Works Association, vol. 94, no. 6, p. 66, Jun. 2002. [Online]. Available: http://www.awwa.org/publications/ journal-awwa/abstract/articleid/14600.aspx [9] American Water Works Association, 3rd ed., ser. AWWA manual. Water audits and loss control programs, Denver, Colorado: American Water Works Asso- ciation, 2009, no. M36. 161 [10] World Health Organization and United Nation’s Children’s Fund Joint Water Supply and Sanitation Monitoring Programme for Water Supply and Sanitation, Progress on drinking water and sanitation: special focus on sanitation. World Health Organization and UNICEF, 2008. [11] V. K. Babbar and Vikram Singh, “Towards rational capital water supply system,” DJB Planning Meeting, 2011. [Online]. Available: http://www.urbanindia.nic. in/programme/uwss/uiww/PPT_4th_Meeting/DJB_Water_PPT.pdf [12] World Health Organization and United Nation’s Children’s Fund Joint Water Supply and Sanitation Monitoring Programme for Water Supply and Sanitation, Global water supply and sanitation assessment 2000 report. World Health Organization and UNICEF, 2000. [13] H. W. King, Handbook of Hydraulics for the Solution of Hydraulic Problems. McGraw-Hill Book Company, Incorporated, 1918. [14] India Central Public Health and Environmental Engineering Organisation, ual on water supply and treatment. Man- Central Public Health and Environmental Engineering Organisation, Ministry of Urban Development, 1999. [15] M. Echenique and R. Seshagiri, “Attribute-based willingness to pay for improved water services: a developing country application,” B: Planning and Design, Environment and Planning vol. 36, no. 3, pp. 384–397, 2009. [Online]. Available: http://dx.doi.org/10.1068/b33126 [16] E. E. Ameyaw, F. A. Memon, intermittent water supply systems,” Technology—AQUA, and J. Bicik, “ Improving equity in Journal of Water Supply: Research and vol. 62, no. 8, p. 552, Dec. 2013. [Online]. Available: http://www.iwaponline.com/jws/062/jws0620552.htm [17] P. F. Boulos, B. W. Karney, D. J. Wood, and S. Lingireddy, “Hydraulic transient guidelines,” Journal of American Water Works Association, pp. 111–124, 2005. [Online]. Available: vol. 97, no. 5, http://www.awwa.org/publications/ journal-awwa/abstract/articleid/15239.aspx [18] D. H. Axworthy and B. W. Karney, “Discussion: undulating elevation profiles,” Filling of pipelines with Journal of Hydraulic Engineering, no. 12, pp. 1170–1174, 1997. [Online]. Available: vol. 123, http://dx.doi.org/10.1061/ (ASCE)0733-9429(1996)122:10(534) [19] Bryan Karney, “Personal correspondence,” Mar. 2013. [20] M.-H. Zérah, Water, unreliable supply in Delhi. Manohar Publishers, 2000. [21] A. C. McIntosh, Asian Development Bank, and International Water Association, Asian water supplies: reaching the urban poor: a guide and sourcebook on urban 162 water supplies in Asia for governments, utilities, consultants, development agencies, and nongovernment organizations. London: Asian Development Bank, 2003. [22] B. E. Evans, pay studies “Willing to pay but unwilling to charge: make a difference?” UNDP, World Do willingness to Bank, and Water and Sanitation Program — South Asia, Tech. Rep., 1999. [Online]. Available: https: //www.wsp.org/sites/wsp.org/files/publications/94200711902_sawilling.pdf [23] S. K. Pattanayak, J.-C. Yang, D. Whittington, and K. C. Bal Kumar, “ Coping with unreliable public water supplies: averting expenditures by households in Kathmandu, Nepal,” Water Resources Research, vol. 41, no. 2, Feb. 2005. [Online]. Available: http://dx.doi.org/10.1029/2003WR002443 [24] M. A. Altaf, “ Household a large secondary city: Habitat International, demand for improved water and sanitation findings from a study in Gujranwala, vol. 18, no. 1, pp. 45–55, in Pakistan,” 1994. [Online]. Available: http://dx.doi.org/10.1016/0197-3975(94)90038-8 [25] P. Klingel, “Von intermittierender zu kontinuierlicher wasserverteilung in entwicklungsländern (from intermittent to continuous water distribution in developing countries),” Ph.D. dissertation, Institute for Water and River Basin Management, Karlsruhe Institute of Technology, Karlsruhe, Germany, 2010. [Online]. Available: http://digbib.ubka.uni-karlsruhe.de/volltexte/1000019357 [26] Legistlative Assembly of the National Capital Territory of Delhi, “ The Delhi Jal Board Act 1998,” Apr. 1998. [Online]. Available: http://ielrc.org/content/ e9802.pdf [27] K. Vairavamoorthy and M. Mansoor, “ Demand management in developing countries,” in Water Demand Management, D. Butler and F. A. Memon, Eds. Lon- don: IWA Publishing, 2006. [28] K. Vairavamoorthy, J. Yan, H. M. Galgale, and S. D. Gorantiwar, “ IRA-WDS: a GIS-based risk analysis tool for water distribution systems,” Modelling & Software, Environmental vol. 22, no. 7, pp. 951–965, 2007. [Online]. Available: http://dx.doi.org/10.1016/j.envsoft.2006.05.027 [29] K. Vairavamoorthy, J. Yan, and S. D. Gorantiwar, “ Modelling the risk of contaminant Management, intrusion vol. 160, in water no. 2, mains,” Proceedings of the ICE - Water pp. 123–132, Jan. 2007. [Online]. Available: http://dx.doi.org/10.1680/wama.2007.160.2.123 [30] R. Franceys and A. Jalakam, “The karnataka urban water sector improvement project; 24x7 water supply is achievable,” Water and Sanitation Program, The World Bank, Tech. Rep., Sep. 2010. [Online]. Available: https://www.wsp.org/ sites/wsp.org/files/publications/WSP_Karnataka-water-supply.pdf 163 [31] US EPA, [32] US Cross-connection Control Manual. EPA, “Methods devices for the prevention Cross-Connection Control Manual, backsiphonage,” in [Online]. and Office of Water, 2003. Available: of backflow and Feb. 2003, pp. 12–16. http://water.epa.gov/infrastructure/drinkingwater/pws/ crossconnectioncontrol/upload/2003_04_09_crossconnection_chapter04.pdf [33] P. L. Skousen, Valve handbook. New York: McGraw-Hill, 1998. [34] Jordan Valve, “Back pressure regulators: Mark 50 self-operated back pressure regulators.” [Online]. Available: http://www.jordanvalve.com/sites/default/ files/mk50.pdf [35] Spirax Sarco, able: “Types of safety valve.” [Online]. Avail- http://www.spiraxsarco.com/resources/steam-engineering-tutorials/ safety-valves/types-of-safety-valve.asp [36] W. A. Conrad, “Pressure-flow relationships in collapsible tubes,” tions on Biomedical Engineering, IEEE Transac- vol. BME-16, no. 4, pp. 284–295, Oct. 1969. [37] F. P. Knowlton and E. H. Starling, “The influence of variations in temperature and blood-pressure on the performance of the isolated mammalian heart,” Journal of Physiology, The vol. 44, no. 3, pp. 206–219, May 1912. [Online]. Available: http://www.ncbi.nlm.nih.gov/pmc/articles/PMC1512817/ [38] C. and D. Bertram flow-induced Structures, and J. Tscherry, oscillations vol. 22, no. 8, in “The collapsible pp. 1029–1045, onset tubes,” of flow-rate limitation Journal of Fluids and Nov. 2006. [Online]. Available: http://dx.doi.org/10.1016/j.jfluidstructs.2006.07.005 [39] R. E. Grandage, “ Development of a variable pressure reducing valve to minimise Papers presented at the International Conference on Developments in Valves & Actuators for Fluid Control, Oxford, England 10-12 September 1985, R. Pickford, Ed. Oxford: BHRA, Sep. 1985, leakage in water distribution systems,” in pp. 251–276. [40] D. McCloy and H. R. Martin, Control of fluid power: analysis and design. New York: E. Horwood, 1980. [41] US EPA, “Commonly asked questions: water act and the NSF standard,” Section 1417 of the safe drinking 2012. [Online]. Available: http: //water.epa.gov/scitech/drinkingwater/dws/plumbing.cfm [42] NSF International Standard and American National Standard, system components - Health effects: NSF/ANSI 61 – 2007a. gan: NSF International, 2007. 164 Drinking water Ann Arbor, Michi- [43] “Parker o-ring handbook (ORD 5700),” Parker Hannifin Corporation, Cleveland, Tech. Rep., 2007. [Online]. Available: http://www.parker.com/literature/ ORD%205700%20Parker_O-Ring_Handbook.pdf [44] E. Oberg, F. D. Jones, H. L. Horton, P. B. Schubert, and G. Garratt, Machinery’s handbook: a reference book for the mechanical engineer, draftsman, toolmaker and machinist, 29th ed., C. McCauley, “O-rings,” Ed. in New York: Industrial Press, 2012, pp. 2587–2592. [Online]. Available: http://cds.cern.ch/record/230198 [45] “National fluid power association (NFPA) diagrams - 62475KAC,” 2010. [Online]. Available: http://www.mcmaster.com/#62475kac/=rzoxie [46] Hydreka, able: “Data logger vista + datasheet.” [Online]. Avail- http://www.hydreka.fr/biblioweb/documents/UK/Data%20sheet/Data% 20Loggers/DOC_Vista+_0710UK.pdf [47] B. E. Larock, R. W. Jeppson, and G. Z. Watters, Hydraulics of pipeline systems. Boca Raton, FL: CRC Press, 2000. [48] G. Ebacher, M. C. Besner, B. Clément, and M. Prévost, “Sensitivity analysis of some critical factors affecting simulated intrusion volumes during a low pressure Research, transient vol. 46, event no. in 13, a full-scale pp. water 4017–4030, distribution Sep. 2012. system,” [Online]. Water Available: http://dx.doi.org/10.1016/j.watres.2012.05.006 [49] J. H. Lienhard (V) and J. H. Lienhard (IV), “Velocity coefficients for free jets from sharp-edged orifices,” Journal of Fluids Engineering, vol. 106, no. 1, pp. 13–17, Mar. 1984. [Online]. Available: http://dx.doi.org/10.1115/1.3242391 [50] R. L. Panton, Incompressible flow, 4th ed. Hoboken, New Jersey: Wiley, 2013. [51] A. Sargaonkar and V. Deshpande, “ Development of an overall index of pollution for surface water based on a general classification scheme in Indian context,” Environmental Monitoring and Assessment, vol. 89, no. 1, pp. 43–67, 2003. [Online]. Available: http://dx.doi.org/10.1023/A:1025886025137 [52] World Health Organization, Guidelines for drinking-water quality, Geneva: World Health Organization, 1997, vol. 3. 165 2nd ed. THIS PAGE INTENTIONALLY LEFT BLANK 166 Appendix A Measuring the precision and accuracy of the pressure loggers The accuracy of the pressure loggers was quoted at for the range 0 to 102m. ±0.1% Best Fit Straight Line A calibration check of their ability to measure negative pressures showed less accuracy and more variability in this range. A.1 Measurement method Using a set of standard gym weights suspended from a hydraulic cylinder a series of known negative pressures were created. These pressures were measured both with 10 Hydreka loggers and a reference calibrated pressure sensor (with an accuracy of ±0.25psi). A box plot for each logger at each pressure point is shown in Figure A-1. Measuring the total variation in the loggers gave the detailed variation data summarized in Table A.1. 99.5% of the measurements between 0 & -4m of pressure were within ±0.45m of the reference pressure. Measurements below -4m inflated the mag- nitude of negative pressure. Since only two data points had pressure in this range: J01 and J02, all measurements below -4m were either taken as -4m so as not to inflate the contamination risk. To test the conclusions made in the results section for their robustness, a MonteCarlo simulation was applied where the variation in the normal error is π= 0.45π . 3 167 168 Pressure (m) Pressure (m) −4.1 −4 −3.9 −3.8 −3.7 −1.7 −1.6 −1.5 −1.4 −1.3 −1.2 −0.4 −0.2 0 0.2 0.4 0.6 1 1 1 2 2 2 3 3 3 4 4 4 5 6 7 Logger Unit Piston + 15lbs 5 6 7 Logger Unit Piston + 5lbs 5 6 7 Logger Unit Atmospheric Pressure 8 8 8 9 9 9 10 10 10 Pressure (m) Pressure (m) Pressure (m) shown as a dashed black line. Pressure (m) Figure A-1: Accuracy and precision or pressure loggers. The reference pressure is −6.2 −6 −5.8 −5.6 −5.4 −5.2 −5 −3.1 −3 −2.9 −2.8 −2.7 −2.6 −2.5 −0.8 −0.6 −0.4 −0.2 1 1 1 2 2 2 3 3 3 5 6 7 Logger Unit 5 6 7 Logger Unit 4 5 6 7 Logger Unit Piston + 20lbs 4 Piston + 10lbs 4 Self−Weight of Piston 8 8 8 9 9 9 10 10 10 Table A.1: Total variation in the logging equipment 169 THIS PAGE INTENTIONALLY LEFT BLANK 170 Appendix B Other ideas for future work In the development and scoping of this project, several areas of supporting knowledge were not explored as fully as they could have been. Future explorations into these would allow for a stronger foundation and likely a more robust design. These include: 1. This thesis used primitive statistical methods to assess the variability, uncertainty, and confidence intervals associated with field-measured pressure and contamination risk. Adding more formal methods would improve the credi- bility and accuracy of the presented conclusions. Future papers published on this data will attempt to incorporate more formal methods of quantifying and bounding the variability and uncertainty. 2. In moving beyond the Beta prototype, the theft risk will be minimized by a design that shifts from metal to plastic and is as lightweight as possible. A more detailed accounting for the predicted loads will be required to achieve such an optimization. Over the course of scoping, executing, and analyzing this project, several divergent paths for alternative projects were identified. These included: 1. Providing a two-tiered storage system at the household level would allow households to gather water during the full duration of supply time, but the lower quality water initially provided by the water company would not be mixed in 171 with the higher quality water that is provided after the system has flushed the contaminants that intruded while it was turned off. 2. As briefly discussed in the background section, the degradation of a distribution system is accelerated by the severe pressure transients associated with filling and draining pipe networks. Quantifying this damage, and exploring options for mitigating it could be an important contribution to the body of knowledge surrounding how to incrementally improve intermittently supplied water. 172 Appendix C Impact of the valve on each house While the results presented in Chapter 6 focused on the aggregated performance of the valve on all houses, for reference, this appendix documents the valve’s impact on each house included in the Results chapter. All nineteen houses that met the requirement that experimental data was gathered for at least 24 hours with the valve and without it were included in Chapter 6 and are documented herein. A single exception was a twentieth house that had played with the bypass valve during the control period; because the pressure at this connection was high, it could not be determined when and for how long the valve had been activated. This connection was, therefore, neither included in Chapter 6 nor in this appendix. 173 Pressure Profile 15 Control Valve Pressure (m) 10 5 0 −5 Avg. Min/Day at this Pressure −10 01/19/14 01/20 01/21 01/22 01/23 Date 01/24 01/25 01/26 01/27 Pressure Histogram 250 Control Valve 200 150 100 50 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 100 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-1: Pressure summary: Azad Market: J01 174 Pressure Profile 15 Control Valve Pressure (m) 10 5 0 −5 Avg. Min/Day at this Pressure −10 01/22/14 01/23 01/24 01/25 01/26 Date 01/27 01/28 01/29 01/30 Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-2: Pressure summary: Azad Market: J02 175 Pressure Profile Pressure (m) 15 Control Valve 10 5 0 −5 01/22/14 01/23 01/24 01/25 01/26 01/27 01/28 01/29 Avg. Min/Day at this Pressure Date Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-3: Pressure summary: Azad Market: J03 176 Pressure Profile Pressure (m) 15 Control Valve 10 5 Avg. Min/Day at this Pressure 0 01/24/14 01/25 01/26 01/27 Date 01/28 01/29 01/30 Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-4: Pressure summary: Azad Market: J04 177 Pressure Profile 6 Control Valve Pressure (m) 4 2 0 −2 Avg. Min/Day at this Pressure −4 01/22/14 01/23 01/24 01/25 01/26 Date 01/27 01/28 01/29 01/30 Pressure Histogram 400 Control Valve 300 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 100 50 0 −50 −100 −150 −10 −5 0 5 Pressure Bin (m) 10 Figure C-5: Pressure summary: Azad Market: J05 178 Pressure Profile Pressure (m) 30 Control Valve 20 10 0 −10 01/23/14 01/24 01/25 01/26 01/27 01/28 01/29 01/30 Avg. Min/Day at this Pressure Date Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-6: Pressure summary: Pitampura: J06 179 Pressure Profile Pressure (m) 20 Control Valve 10 0 Avg. Min/Day at this Pressure −10 01/20/14 01/21 01/22 01/23 01/24 Date 01/25 01/26 01/27 01/28 Pressure Histogram 500 Control Valve 400 300 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-7: Pressure summary: Pitampura: J07 180 Pressure Profile Pressure (m) 10 Control Valve 5 0 Avg. Min/Day at this Pressure −5 01/25/14 01/26 01/27 01/28 Date Pressure Histogram 01/29 01/30 1500 Control Valve 1000 500 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 1000 500 0 −500 −1000 −10 −5 0 5 Pressure Bin (m) 10 Figure C-8: Pressure summary: Vivek Vihar: J08 181 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 Avg. Min/Day at this Pressure −5 01/16/14 01/17 01/18 01/19 01/20 01/21 01/22 Date 01/23 01/24 01/25 01/26 01/27 Pressure Histogram 400 Control Valve 300 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 100 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-9: Pressure summary: Pitampura: J09 182 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 −5 01/16/14 01/17 01/18 01/19 01/20 01/21 01/22 01/23 Avg. Min/Day at this Pressure Date Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 150 100 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-10: Pressure summary: Pitampura: J10 183 Pressure Profile Pressure (m) 15 10 5 0 −5 01/16/14 Avg. Min/Day at this Pressure Control Valve 01/17 01/18 01/19 01/20 Date 01/21 01/22 01/23 01/24 Pressure Histogram 400 Control Valve 300 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 300 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-11: Pressure summary: Pitampura: J11 184 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 Avg. Min/Day at this Pressure −5 01/16/14 01/17 01/18 01/19 01/20 Date 01/21 01/22 01/23 01/24 Pressure Histogram 400 Control Valve 300 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 100 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-12: Pressure summary: Pitampura: J12 185 Pressure Profile Pressure (m) 10 Control Valve 5 0 Avg. Min/Day at this Pressure −5 03/31/14 04/01 04/02 04/03 04/04 04/05 Date 04/06 04/07 04/08 04/09 Pressure Histogram 300 Control Valve 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-13: Pressure summary: Vasant Vihar: M01 186 Pressure Profile Pressure (m) 15 10 5 0 −5 03/31/14 Avg. Min/Day at this Pressure Control Valve 04/01 04/02 04/03 04/04 04/05 Date Pressure Histogram 04/06 04/07 04/08 04/09 500 Control Valve 400 300 200 100 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 100 50 0 −50 −10 −5 0 5 Pressure Bin (m) 10 Figure C-14: Pressure summary: Vasant Vihar: M02 187 Pressure Profile Pressure (m) 10 Control Valve 5 0 −5 04/02/14 04/03 04/04 04/05 04/06 04/07 04/08 04/09 Avg. Min/Day at this Pressure Date Pressure Histogram 800 Control Valve 600 400 200 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 0 −200 −400 −10 −5 0 5 Pressure Bin (m) 10 Figure C-15: Pressure summary: Shivalik: M03 188 Pressure Profile Pressure (m) 4 Control Valve 2 0 −2 −4 04/02/14 04/03 04/04 04/05 04/06 04/07 04/08 04/09 Avg. Min/Day at this Pressure Date Pressure Histogram 800 Control Valve 600 400 200 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 0 −200 −400 −10 −5 0 5 Pressure Bin (m) 10 Figure C-16: Pressure summary: Shivalik: M04 189 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 Avg. Min/Day at this Pressure −5 03/30/14 03/31 04/01 04/02 04/03 04/04 Date 04/05 04/06 04/07 04/08 04/09 Pressure Histogram 1500 Control Valve 1000 500 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 200 100 0 −100 −200 −10 −5 0 5 Pressure Bin (m) 10 Figure C-17: Pressure summary: Lado Sarai: M05 190 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 Avg. Min/Day at this Pressure −5 03/30/14 03/31 04/01 04/02 04/03 04/04 Date 04/05 04/06 04/07 04/08 04/09 Pressure Histogram 1500 Control Valve 1000 500 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 150 100 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-18: Pressure summary: Lado Sarai: M06 191 Pressure Profile 20 Control Valve Pressure (m) 15 10 5 0 Avg. Min/Day at this Pressure −5 03/30/14 03/31 04/01 04/02 04/03 04/04 Date 04/05 04/06 04/07 04/08 04/09 Pressure Histogram 1500 Control Valve 1000 500 0 −10 −5 0 5 10 Pressure Bin (m) Avoided Pressure Histogram (Control−Valve) 15 20 15 20 Avg Min/Day Avoided 100 50 0 −50 −100 −10 −5 0 5 Pressure Bin (m) 10 Figure C-19: Pressure summary: Lado Sarai: M07 192