You are on page 1of 71

Samoa PRO Turbine & Generator Design

Subsection 6

Scott Clowser

Authors Rowan Cooper-Caroselli

Scott Ruud

Ian Snyder

ENGR 492 Fall 2013

Table of Contents
List of Figures ....................................................................................................................... iii List of Tables ........................................................................................................................ iv Executive Summary ............................................................................................................... v Introduction .......................................................................................................................... 1 Literature Review Hydroturbine and Generator .................................................................. 1
Hydroturbines ................................................................................................................................1 Hydropower Theory .............................................................................................................................. 3 Concept of Specific Speed ..................................................................................................................... 3 Reaction Turbines ..........................................................................................................................5 Francis Turbine ...................................................................................................................................... 5 Impulse Turbines ..........................................................................................................................11 Pelton Wheel....................................................................................................................................... 11 Turgo Turbine...................................................................................................................................... 14 Ossberger (Cross-Flow) Turbine.......................................................................................................... 15 General Considerations ................................................................................................................17 Turbine Erosion Consideration ......................................................................................................19 Materials for Seawater Turbines ........................................................................................................ 20 Case Study: Materials Selection for a Japanese Pilot Seawater Pumped Storage Power Plant ....... 21 Material Hard Coatings for Resistance to Hydro-abrasion ................................................................. 22 Cavitation ............................................................................................................................................ 22 Alternating Current (AC) and Power Factor (PF).............................................................................24 Electro Mechanical Equipment .....................................................................................................25 Generators .......................................................................................................................................... 25 Power Quality and Conditioning ...................................................................................................29

Operation and Maintenance ................................................................................................ 30 Regulatory Framework ........................................................................................................ 31


Federal Regulatory Agencies .........................................................................................................31 State Regulatory Agencies ............................................................................................................31 Relevant State Laws and Decisions ..................................................................................................... 31 Local Regulatory Agency ...............................................................................................................32 Grid Connection and Virtual Net Metering ....................................................................................32

Power Sales ......................................................................................................................... 33 Turbine Component Design ................................................................................................. 34


Constraints and Assumptions .......................................................................................................34 Selection Criteria ..........................................................................................................................35 i

Turbine Design Alternatives ..........................................................................................................35 Pelton Wheel Design ........................................................................................................................... 35 Turgo Runner Design........................................................................................................................... 37 Crossflow Turbine Design ................................................................................................................... 38 Francis Turbine Design ........................................................................................................................ 40 Preferred Alternative ...................................................................................................................41 Comparison of Alternatives with Selection Criteria............................................................................ 41 Choice of Preferred Alternative .......................................................................................................... 41 Selection of System Generator ........................................................................................................... 42 Design Detail ................................................................................................................................43

Economics of the PRO Power Plant ...................................................................................... 46 Conclusions ......................................................................................................................... 48 Works Cited......................................................................................................................... 49 Appendix............................................................................................................................. 53

ii

List of Figures
Figure 1: Operating Ranges for Main Hydroturbine Types (Alternate Hydro Energy Center, 2011; Andritz Hydro, 2013; Andritz Hydro, 2013; Andritz Hydro, 2013) ..........................................................................................................1 Figure 2: Micro and Small Hydro Turbine Operating Ranges (Paish, 2002) ...................................................................2 Figure 3: Turbine Operating Ranges (Gilkes Hydropower, 2013) ..................................................................................2 Figure 4: Efficiency vs. Specific Speed for Primary Turbine Types (Dixon & Hall, 2010) .................................................4 Figure 5: Simple vertical shaft Francis turbine schematic (VDE, 2008) ..........................................................................5 Figure 6: Diagram of a Francis Turbine showing Guide Vanes and Runner (Boyle, 2004) .............................................6 Figure 7: Sectional Sketch of Francis Turbine Blading Showing Velocity Diagrams at Runner Inlet and Exit (Dixon 2002) ..............................................................................................................................................................................7 Figure 8: Comparison of Velocity Triangles for a Francis Turbine at Full and Partial Load Operation (Dixon, 2010) ....8 Figure 9: Hydraulically Controlled Piston for Adjusting Guide Vane Angle (in yellow) (Microhydro, 2013) ..................8 Figure 10: Location of Draft Tube in Relation to Vertical Shaft Francis Turbine (Dixon 2010). .....................................9 Figure 11: Canyon Hydro 910 KW Francis Turbine and Generator (Canyon Hydro, 2013) ..........................................11 Figure 12: Schematic of a Pelton Wheel Turbine (Boyle, 2004) ...................................................................................12 Figure 13: Jet Striking Runner (Meazza, 1996) ............................................................................................................12 Figure 14: Simplified Turgo Jet and Runner (European Small Hydro Association, 1998) .............................................14 Figure 15: Turgo Jet Entry and Exit Angles (Crewdsonem, 2013) ................................................................................15 Figure 16 : OSSBERGER vertical inflow schematic (OSSBERGER, 2013) .......................................................................16 Figure 17: Diagram of an Ossberger (Crossflow) Turbine (OSSBERGER, 2013)............................................................16 Figure 18: Part-flow Efficiency Curves for Different Turbines (European Small Hydro Association, 2010) ..................18 Figure 19: Speed Increasers Parallel shaft (top left), Bevel (top right), and Belt drive (bottom) (European Small Hydropower Association, 2004) ...................................................................................................................................19 Figure 20: Cavitation versus specific speed used to determine degree of cavitation ..................................................23 Figure 21: Three Phase Synchronous Generator (PDH, 2012) ......................................................................................26 Figure 22: Torque and Speed characteristics of an induction machine (Heng, 1992). .................................................27 Figure 23: Selection of a Generator Informed by the Operating Speed (of which the number of poles is a function) and the Power Output Available from the Prime Mover (ABB, 2011). ........................................................................28 Figure 24: Example system with selected design characteristics (Hydrolink, 2009) ....................................................43 Figure 25: Proposed general site layout ......................................................................................................................44 Figure 26: Simple diagram of the horizontal Francis turbine and generator setup .....................................................45 Figure 27: Aggregate capital costs for the PRO power plant.......................................................................................46 Figure 28: Aggregate O&M costs for the PRO power plant .........................................................................................47 Figure 29: Net power determination is based on parasitic loading in each plant sub-system; two analyses show the variations in design conclusions from the freshwater 1 and freshwater 2 groups. .....................................................47

iii

List of Tables
Table 1: Corrosion Rates of Common Materials Used for Seawater Pump Applications (NiDI, 1995) .........................21 Table 2: Relative Cost Factors of Various Materials for Seawater Piping Application (Copper Development Association, 1986) .......................................................................................................................................................21 Table 3 : Performance of Various Hard Coatings .........................................................................................................22 Table 4: Common causes of power quality problems (Brown, 2006) ..........................................................................30 Table 5: Design constraints for selection of the hydroturbine. ....................................................................................34 Table 6: Pelton Turbine Design Parameters.................................................................................................................36 Table 7: Pelton Design Outputs ...................................................................................................................................36 Table 8: Advantages and Disadvantages of the Pelton Wheel ....................................................................................37 Table 9: Turgo Turbine Design Parameters .................................................................................................................37 Table 10: Turgo Design Outputs ..................................................................................................................................38 Table 11 : The advantages and disadvantages of the Turgo Runner...........................................................................38 Table 12: Crossflow Turbine Design Inputs ..................................................................................................................39 Table 13: Crossflow Turbine Design Outputs ...............................................................................................................39 Table 14: Advantages and disadvantages of the cross flow turbine ...........................................................................40 Table 15: Francis Turbine Design Specifications ..........................................................................................................40 Table 16: Summary Table of Alternatives and Selection Criteria .................................................................................41 Table 17: Comparison of Pumping Cost using Pelton Wheel vs. Residual Head using Francis Turbine .......................41 Table 18: Generator Specifications ..............................................................................................................................42 Table 19: Parameters of final design ...........................................................................................................................44

iv

Executive Summary
Although pressure retarded osmosis is an emerging technology, the application of a hydro-turbine within the PRO system is essentially the same as in classical hydropower systems. Instead of using a reservoir with a penstock or an in-stream/run-of-river turbine, a portion of the pressurized drawsolution flow is piped through the turbine. The rate of flow provided to the turbine will be equal to the fresh water flux across the PRO membranes. Like any other hydro resource, the specific turbine setup is dictated primarily by available head and flow characteristics. Four turbine selection alternatives were compared to determine the best design selection. The Francis Turbine was chosen as the preferred alternative based on the following qualifications: Best lifetime economics of any alternative Highest efficiency Highest lifetime net revenue Eliminates the need for an extra pump to move water to wet well

Paired with the Francis turbine is a three-phase induction generator. The reason for selecting an induction generator is motivated by the low capital cost and O&M cost associated with this option coupled with the lack of operational requirements that might justify upgrading to a synchronous machine. The plant generates 230kW 285kW of net power, which equates to a first year net revenue of $189,000 - $234,000 (future changes in energy price will affect this revenue). The capital cost of the entire PRO system is estimated at $31 to $35 million with an annual operations and maintenance cost of $2.3 million to $2.4 million. The capital cost of the turbine generator subsystem is estimated at $2.2 million, roughly 3% of the overall capital cost of the system. The annual operation and maintenance cost of the turbine/generator subsystem is estimated at $44,000. The net present value of the entire PRO energy generation facility is estimated to be -$60 million. The key findings of the preliminary analysis are: Commercial scale project is not economically feasible at this site o Freshwater intakes require significant energy o Freshwater treatment is costly PRO is a dilute energy source o Seawater intake and treatment require significant energy o Energy losses through membrane module are significant More research on membrane fouling and membrane module design is necessary to accurately predict water treatment energy requirements Good potential site for pilot scale research facility

Introduction
Although pressure retarded osmosis is an emerging technology, the application of a hydroturbine within the PRO system is essentially the same as in classical hydropower dam scenarios. Instead of using a reservoir with a penstock or an in-stream/run-of-river turbine, a portion of the pressurized drawsolution flow is piped through the turbine. The rate of flow provided to the turbine will be equal to the fresh water flux across the PRO membranes. Like any other hydro resource, the specific turbine setup is dictated primarily by available head and flow characteristics.

Literature Review Hydroturbine and Generator


Hydroturbines
There are many hydroturbine options available for power production, but the exact choice of turbine and achievable efficiency is a function of the operating pressure and flow of the final design. Impulse type turbines (Pelton, Ossberger and Turgo) are typical for moderate to high head applications. Reaction type turbines (Francis, Kaplan, and Propeller) are used for low to moderate head applications and generally higher flow (Paish, 2002). Operating ranges for various types of turbines for smaller-scale hydro are often displayed graphically (Figure 1 & Figure 2).

Figure 1: Operating Ranges for Main Hydroturbine Types (Alternate Hydro Energy Center, 2011; Andritz Hydro, 2013; Andritz Hydro, 2013; Andritz Hydro, 2013)

Figure 2: Micro and Small Hydro Turbine Operating Ranges (Paish, 2002)

The reported operating ranges differ slightly among sources, but it is certain that the expected net head available for a full-scale PRO facility will exceed the typical range of head for Kaplan and propeller turbines at the expected design flow. These turbine types have therefore been dismissed as viable options for the facility. The most likely turbines that have operating ranges encompassing the expected facility operating range are the Turgo runner, Pelton wheel, and Francis turbine. The operating ranges for head and flow on these turbines give generation potentials of 20 kW to 20 MW are (Figure 3). The Ossberger (cross-flow) turbine is also a potential design option.

Figure 3: Turbine Operating Ranges (Gilkes Hydropower, 2013)

Hydropower Theory The governing equation that quantifies the conversion of either the potential or kinetic energy of water into power, assuming no losses is: = (1)

In the above equation, the variable is in Watts, is the density of water in kg/m3, is the constant of gravitational acceleration in m/s2, is the design flow in m3/s, and is the design head in meters. Incorporating overall turbine and piping inefficiency as and including frictional loss in the head, the equation reduces to: = (2)

Depending on the type of turbine, the value of will vary between 0.65 and 0.95. The head,, can be calculated by finding the friction factor over the length of the penstock and the minor losses at junctions or bends. The pipe friction loss can be calculated using the Darcy-Weisbach equation: = 2 2 (3)

Where is the pipe head loss from friction, is the Darcy friction factor, is the length of the pipe, is the diameter of the pipe, is the velocity of the moving water, and is the constant of gravitational acceleration. The Darcy friction factor for turbulent flow is commonly calculated by using a root-finder on the following relationship: / 2.51 2 log + = 0 3.7 1 (4)

Where is the Darcy friction factor, is a roughness coefficient, is the diameter of the pipe, and is the Reynolds number. Minor losses through the pipe are typically incorporated using a loss factor ( ) which depends on the geometry of the valve, bend, or other component. These minor head losses are proportional to the loss factor and the square of velocity of the working fluid (Eq. 5). The other variables are defined in the same way as previously outlined. = 2 2 (5)

Once the net head has been estimated for a site, the turbine power output can be estimated using typical efficiencies, as in Equation 2. The output to the grid is then found by incorporating the generator and drive system efficiency. The efficiency of a turbine is dependent on the type of turbine used, the sophistication of the engineering and design, and how well matched the turbine is to the site. Concept of Specific Speed One useful quantity used in turbine design is the specific speed, typically denoted as . The specific speed of a turbine is a pseudo-dimensionless value that describes the speed of a unit that produces one 3

unit output for one unit head. Appropriate unit cancelling using fundamental constants would be required to make it truly dimensionless, but in practice the specific speed is defined in SI units as: = ( )1.25 (6)

The units of specific speed are an exponentiation of kW, rpm, and meters. The variable is the power in kW, and is the rotation in rpm. This form of the specific speed equation has become fairly standardized for turbine selection (Warnick, 1984). Turbines are often compared based on their specific speed. The specific speed is typically calculated and reported at the point of peak efficiency for a turbine, although it can be calculated for any value (Warnick, 1984). An alternate and truly unit-less form of the specific speed is also often used for sizing turbines, typically denoted sp and referred to as the power specific speed. The power specific speed for a given turbine can be calculated as follows: = ( )4
5

(7)

The variable is the rotational speed in radians per second and the other variables remain as defined previously. The numerical values of the two specific speed equations for identical inputs varies greatly, therefore they must be carefully applied to known charts or tables describing the optimal operating ranges for a given turbine type. Using the dimensionless version, the operating ranges for the primary turbine types and possible maximum efficiencies are shown in Figure 4.

Figure 4: Efficiency vs. Specific Speed for Primary Turbine Types (Dixon & Hall, 2010)

Reaction Turbines
Reaction turbines operate completely housed and immersed, generating power from the torque applied to the turbine runner from the water pressure. There is a pressure difference across the turbine, and the housing must be able to withstand the pressure of the applied head at the inlet to the turbine. The primary types of reaction turbines include the Francis turbine, Kaplan turbine, and other propeller turbines. However, due to the flow and pressure ranges for the PRO process, only the Francis turbine is considered applicable for design. A simple diagram of a Francis turbine is shown in Figure 5.

Figure 5: Simple vertical shaft Francis turbine schematic (VDE, 2008)

Francis Turbine Francis turbines are designed with a spiral casing called a volute or scroll to guide the water to enter the turbine radially. Water passes through the vanes and exits axially. Francis turbines are typically installed with a vertical axis of rotation and downward vertical exit shaft. This configuration saves space and is typically more cost effective for larger installations (Alternate Hydro Energy Center, 2011). The runner vanes of a Francis turbine are fixed, and flow through the turbine is regulated by a wicket-gate system utilizing guide vanes. In the runner, angular momentum of the working fluid is reduced and work is supplied to the turbine shaft (Dixon & Hall, 2010). A simple diagram is shown in Figure 6.

Figure 6: Diagram of a Francis Turbine showing Guide Vanes and Runner (Boyle, 2004)

The water flows from the runner into a draft tube before entering the tailrace. If the eventual outfall of the system requires a pressure to pump the water, the draft tube geometry can be designed to allow for residual head. This will reduce the efficiency of the turbine, but will avoid the efficiency losses of using a pump after the turbine. At the inlet to the guide vanes, the absolute flow angle 1 is defined as Where 1 is the absolute flow angle in the radial/tangential plane and 1 is the absolute velocity of the water at the inlet of the guide vanes. The absolute velocity, 1 , is defined as magnitude of vectors 1 and 1 (Eq. 8). Where swirl velocity and the radial velocity are signified by and respectively. The flow is guided to angle 2 and velocity 2 , the absolute condition of the flow before entering the runner. By vector subtraction, the relative velocity at the entry to the runner 2 can be found using the following equation: (10) 2 = 2 2 Where 2 and 2 are the absolute and tangential vector velocities of the water within the scroll, shown in Figure 7.
2 2 + 1 1 = 1

1 = tan1

1 1

(8)

(9)

Figure 7: Sectional Sketch of Francis Turbine Blading Showing Velocity Diagrams at Runner Inlet and Exit (Dixon 2002)

The relative flow angle 2 at the inlet to the runner is defined as:

The water leaves the runner with a relative flow angle 3 , a relative flow velocity 3. The absolute velocity at the runner exit is found by vector addition 3 = 3 + 3 . The relative flow angle 3 at the outlet of the runner is defined as: 3 = tan1[(3 + 3 )/3 ] (12)

2 = tan1[(2 2 )/2 ]

(11)

Where 3 and 3 are the swirl and radial velocity at the exit of the runner and 3 is the tangential velocity at the exit of the runner. The equation (above) assumes a swirl velocity 3 is present at the runner exit, but for simple analyses of the Francis turbine, it is assumed that 3 equals zero (Dixon & Hall, 2010).

Operation at Partial Load The power output of the Francis turbine may be reduced by swiveling the guide vanes to restrict the flow of water through the turbine ( ), while maintaining a constant blade velocity, shown in Figure 8.

Figure 8: Comparison of Velocity Triangles for a Francis Turbine at Full and Partial Load Operation (Dixon, 2010)

The efficiency of the turbine changes as a function of the load ratio. The load ratio is reduced from the design flow by adjusting the guide vanes to reduce the flow of water through the turbine. Francis turbines maintain a relatively high efficiency between 50% and 100% of design flow, but they unable to achieve a high efficiencies over a large range of flow. Wicket-Gate controls have been developed to adjust the angle of all the guide vanes simultaneously, as shown in Figure 9. These gate systems can be easily integrated into an overall facility control system to change the amount of power produced based on grid requirements.

Figure 9: Hydraulically Controlled Piston for Adjusting Guide Vane Angle (in yellow) (Microhydro, 2013)

Eulers turbine equation is used to define the power produced from the flow of water through the turbine, resulting in: = 2 2 3 3 = 2 2 (13)

If the flow at runner exit has no swirl then the equation is reduced to: (14)

The energy available at the entry to the runner is equal to the kinetic, potential and pressure energies combine as: ( ) = 2 1 2 + 2 + 2 2 (15)

At the outlet of the runner, the energy in the water has been reduced by the amount of specific work and by friction work in the runner, , and the remaining energy is equal to the sum of the pressure potential and kinetic energies, given by: 1 2 3 + + 3 ( ) = 3 2
(02 03 )

Where is the head loss in the volute and guide vanes, 2 is the absolute static pressure at the runner inlet, and is absolute atmospheric pressure.

(16)

Where 3 is the absolute static pressure at the runner exit. By taking the difference of equations 14 and 15, the equation for specific work is defined as: = + (2 3 ) (17)

Where 02 and 03 are the absolute total pressures at runner inlet and exit, and 3 is equivalent to Z as shown in Figure 10.

Figure 10: Location of Draft Tube in Relation to Vertical Shaft Francis Turbine (Dixon 2010).

The energy equation between the runner exit and the tailrace can now be defined (Eq. 17). 1 2 3 1 2 + 3 + 3 = 4 + 2 2 2 2 3 3 = 2 2 = (18)

Where is the head loss in the draft tube and 4 is the flow exit velocity. The hydraulic efficiency is expressed by: = (19)

Finally, if the flow exiting the runner has no swirl, the equation reduces to: = (20)

The overall efficiency is then defined as:

And the ratio of runner tip speed to the jet velocity is defined as: = 2 1 (22)

(21)

Francis turbines are typically used in conjunction with a synchronous generator and the rotational speeds chosen will provide either 50 or 60 cycles per second. The speed of the turbine must remain constant for this application. Some turbine manufacturers provide the entire turbine-generator system as a package, where the generator provided is engineered to match the turbine output as optimally as possible. The setup provided for a small hydro system package manufactured by Canyon Hydro is shown in Figure 11. The full operating range of a Francis turbine is generally between 40% and 110% of rated design discharge, but the upper limit is flexible depending on the maximum generator rating (Alternate Hydro Energy Center, 2011). The head operating range is approximately between 65% and 125% of design head (Alternate Hydro Energy Center, 2011).

This value has a fairly wide range (0.6 0.95) that still allows for high efficiency.

10

Figure 11: Canyon Hydro 910 KW Francis Turbine and Generator (Canyon Hydro, 2013)

Impulse Turbines
Unlike reaction turbines, impulse turbines do not require a pressurized, sealed housing and are operated at atmospheric pressure. In an impulse turbine the net head is converted into kinetic energy as the water is forced through a regulated nozzle, and the force of this jet impacts the turbine runner. Types of impulse turbines include the Pelton wheel, Turgo runner, and the patented Ossberger (cross-flow). Impulse turbines typically operate at a higher head and lower flow than reaction turbines and therefore extract more energy per volume of water (European Small Hydro Association, 1998). Pelton Wheel The typical operating range for a Pelton wheel turbine is between 20 meters and 1300 meters of hydraulic head, but they are typically utilized for heads higher than 250 meters (Boyle, 2004). Largescale commercial hydropower plants utilizing Pelton turbines typically require a head higher than 500 meters. The general design of a Pelton turbine includes at least one nozzle and one runner (Figure 12), although multiple runners or nozzles are often used. The nozzle converts the energy of the highpressure water into a jet of water with high kinetic energy that impacts the turbine runner. Various literature sources indicate that energy loss due to nozzle friction is typically between 1% and 8%. The buckets of the Pelton runner are designed to redirect the jet of water nearly 180 (typically 165) from the direction of impact (Figure 13). Although a 180 exit angle is the most ideal for maximum energy transfer, the exit angle must slightly less than 180 to ensure that water exiting the bucket does not interfere with other buckets (Warnick, 1984). The smooth curve of the bucket allows the momentum of the water from the jet to be transferred to the runner with minimal losses.

11

Figure 12: Schematic of a Pelton Wheel Turbine (Boyle, 2004)

Figure 13: Jet Striking Runner (Meazza, 1996)

The working theory behind the Pelton wheel begins with the kinetics of an impulse runner, in which torque applied to the axel of the turbine is supplied by the force of one or more jets of high velocity water striking the runner (Eq. 22). = 1 (1 cos ) (23)

Where is the force of the jet on the runner vane in N, is the density of water, is the volumetric flow of water striking the runner vane in m3/s, 1 is the relative velocity of the jet with respect to the moving vane in m/s, is a loss coefficient for movement across vane, and is the deflection angle relative to the original jet (Warnick, 1984). A typical value for is 0.96. The torque exerted by this force is the product of the force of the jet, , and the lever arm, , which is also the radius of the turbine: The theoretical shaft power of the turbine is the product of the force acting on the runner and the velocity of the runner, (m/s): = (25) = (24)

12

The power output, , is in Watts. The relative velocity of the water striking the vane of the runner, 1 , is the difference between the absolute velocity of the jet, , and the absolute linear velocity of the runner, , at moment of impact: All velocities are in m/s. The water exits the runner vane at the deflection angle and experiences velocity loss across the vane, as described previously. This exit velocity, 2 , has no tangential velocity component at optimum turbine speed, so it balances exactly the exiting velocity of the jet along the axis of the incoming jet: 2 = 1 (27) (28) 1 = (26)

This equation can be solved for the runners linear speed, , giving: = cos cos 1

= 2 cos = ( ) cos

(29)

Substituting all of this back into the theoretical power equation yields: =
2 1 cos ( )()(1 cos ) = cos 1

(30)

The jet velocity of the turbine can be calculated by:

Where is denotes the nozzle loss term (typically 0.96 0.99), is the constant of gravitational acceleration in m/s2, and is the net head in m (Warnick, 1984). The jet diameter equation begins with the jet discharge relationship: = 2 (32)

= 2

(31)

Where and are the jet discharge rate and jet cross-sectional area, respectively. By incorporating the jet diameter into the jet discharge equation, a final equation for the jet diameter, , can be developed. Using a typical nozzle loss coefficient of 0.98 for , the relation can be solved by: 4 1/2 0.402 1/4 = 2 (33)

A general width, length, and depth of the Pelton buckets can be calculated as a direct function of the jet diameter. A function for the number of buckets on the runner, , can be found in the literature and is given by: 13

Where is the runner diameter (Nasir, 2013). The runner diameter itself is calculated by the required rotational speed for the generator and calculated linear velocity of the runner. The diameter also must have a minimum ratio to the bucket size so that splash back from each bucket does not interfere with the oncoming bucket. One quick method used is the ratio of the diameter to bucket, which must be greater than 2.7 for efficient operation, and it should preferably be higher (Nasir, 2013). The bucket dimensions themselves, in a theoretical design, are given by: = 3.4 = 1.2 = 3.0 (35) (36) (37)

= 15 + 2

(34)

Turgo Turbine Turgo turbines operate in a range between 15 meters to 300 meters of head (Alternate Hydro Energy Center, 2011). They are very similar to the design of a Pelton wheel, where the main exception is the impact angle of the jet on the runner (Figure 14). Instead of impacting the runner exactly tangentially, as is the case with a Pelton runner, a Turgo jet impacts the runner from outside the plane of the runner. The typical impact angle is 20. This modification results in minimal interference of the exiting water with either the runner blades or the jet. Some of the drawbacks to using a Turgo runner are the slight loss of efficiency and more sophisticated runner design.

Where , , and are the bucket width, length, and depth, respectively (Nasir, 2013)..

Figure 14: Simplified Turgo Jet and Runner (European Small Hydro Association, 1998)

Much of the theory behind the Turgo turbine is either identical or directly analogous to the Pelton wheel theory, and most will not be repeated in this section. The total deflection angle of the jet is the sum of the entry angle of the water (typically 20) and the exit angle (Figure 15).

14

Figure 15: Turgo Jet Entry and Exit Angles (Crewdsonem, 2013)

The calculation of the force and torque on the turbine runner is performed the same way, and the nozzle dimensioning and flow calculations are identical. The system can be sized based on the necessary generator rpm (within the turbine operating range). Because splash back is not a concern, the runner diameter can be much smaller for similar flow. The diameter of the Turgo runner is found by: = 60 (36)

Where is the runner linear velocity and is the rotational speed in rpm (Anagnostopoulos & Papantonis, 2008). After including the loss coefficients in the runner and jet design calculations (analogous to the Pelton equations), the final efficiency is simply found by the design output over the theoretical power potential: = (37)

The vane design and dimensions for a Turgo runner are proprietary in nature, so finding a preliminary method for calculating a design is not available. The Turgo turbine itself was invented by the Gilkes Corporation in 1919, so most of its installations and available data belong to that company and other manufacturers who have since copied the design. Ossberger (Cross-Flow) Turbine The Ossberger turbine is a slow speed radial flow turbine. Guide vanes direct a rectangular water jet through the blade ring of a cylindrical rotor (OSSBERGER, 2013). The water flows across the vanes to the interior of the cylinder and flows out the vanes toward the other side before exiting the turbine casing into the draft tube (Figure 16). The overall efficiency of the OSSBERGER turbine is about 80% for small power outputs over the entire operating range, but efficiencies of 86% have been measured for larger units (OSSBERGER, 2013). Ossberger turbines of lower efficiency and simpler construction are common in hydro schemes in developing countries, but the relatively recent advances in efficiency for wellengineered designs made the Ossberger design a viable option for consideration for the PRO facility. The operating range of cross flow turbines is typically small, but can be operated using up to 200 meters in 15

head and flows of up to 13 m3/s (Free Flow Hydro, 2010). The capital cost for cross-flow turbines can be as low as 50% of that of other turbine types (Aziz & Desai, 1991).

Figure 16 : OSSBERGER vertical inflow schematic (OSSBERGER, 2013)

The guide vane of an Ossberger turbine controls the attack angle and flow through the runner. The installation of the guide vane and housing can either be vertical (Figure 16 and Figure 17) or horizontal. Some systems have an additional guide mechanism that partitions the inlet pipe to some fraction of its full capacity, allowing the system to run efficiently at partial flow.

Figure 17: Diagram of an Ossberger (Crossflow) Turbine (OSSBERGER, 2013)

As another type of impulse turbine, the theory behind the cross-flow turbine and runner is very similar to the Pelton and Turgo theory. The water jet equation is the same, although in the case of the crossflow turbine, it is more like a sheet of water entering the runner rather than a cylindrical jet. The runner tangential velocity can be found by: 1 = 0.5 cos (38) 16

Where 1 is the runner linear velocity, is the jet velocity, and is the attack angle (Zia, Ghani, Wasif, & Hamid, 2010). A very common value for is 22. Finding the optimum runner tangential velocity allows calculation of the runner outer diameter using: 1 = 1 2 (39)

Where 1 is the runner outer diameter and is the runner radial velocity (Zia, Ghani, Wasif, & Hamid, 2010).Some designs specify the runner outer diameter as a design variable, and calculate the opposite direction, using experimentally determined optimal design ratios. The inner diameter can be calculated from a suggested optimum ratio of 0.7 (Aziz & Desai, 1991). The length of the cross-flow runner can be determined by: 1 = 210 (40)

Where is the runner length, is the net head, is the design flow, and 1 is the diameter solved from the previous equation (Zia, Ghani, Wasif, & Hamid, 2010). The number of runner blades is a point of contention among some of the sources, ranging anywhere from 15 to 37 (Aziz & Desai, 1991) (Zia, Ghani, Wasif, & Hamid, 2010). Either the number of blades is calculated for a given optimum spacing, or the number of blades is chosen based on experimental data for a similar sized turbine. An optimum ratio of 1.03 for the runner width to blade spacing was cited by one study on cross-flow design parameters for a 90 exit angle (Aziz & Desai, 1991). The runner width can be determined once the blade radius of curvature, , is known, which can be determined by: The blade exit angle for perfect radial flow should be 90. Many systems are designed using this parameter, but it is suggested that this angle be increased at times to prevent shock losses (Zia, Ghani, Wasif, & Hamid, 2010). The angle between the periphery of the runner blade and the angle of attack can be found using: 1 = tan1 (2 tan 1 ) (42) = 0.326 1 2 (41)

Ultimately, final engineering of the system by a turbine manufacturer will be required. The above analysis only provides a theoretical and preliminary design for a cross-flow turbine.

General Considerations
Another necessary consideration is whether the flow or head through the turbine should or can be easily varied without serious loss of efficiency. The adaptability of impulse-type turbines to varying flow conditions is generally greater than that of reaction-type turbines (Paish, 2002). Figure 18 illustrates the turbine efficiencies for flows that are less than the design flow for various turbines (European Small Hydro Association, 2010).

17

Figure 18: Part-flow Efficiency Curves for Different Turbines (European Small Hydro Association, 2010)

A system specific consideration is the potential need for residual head at the turbine outlet. Impulse turbines generally exit to atmospheric pressure as all the head is converted to velocity that impacts the runner. All the head is lost once the water leaves the runner. Reaction turbines have some ability to retain head after the runner because they operate in a closed housing. For example, the Francis turbine can retain head by designing the draft tube in such a way to induce swirl after the runner, which is a means of retaining some of the energy in the water. This is typically not desirable for most applications, but the unique situation of the project may require head to be retained to ensure sufficient effluent flow. Some turbines will also require speed increasers between the turbine shaft and the generator shaft. Hydro turbines sometimes rotate at speeds that are slower than what is required for a typical alternator or synchronous generator, so a device to increase the rotational speed must be used (European Small Hydropower Association, 2004). This is more typical of impulse turbines, but any turbine may require a speed increaser, depending on the net head and flow. Some typical types of speed increasers include parallel shaft gearing, bevel gearing, and belt drives (European Small Hydropower Association, 2004) (Figure 18). Belt drives generally have a maximum speed ratio of 3:1 (Harvey, 1993). No matter what type of speed increaser is used, some efficiency loss will occur. Friction losses can also be a substantial issue, especially in poorly maintained systems, resulting in power losses up to 8% of the gross shaft power (European Small Hydropower Association, 2004). There is also an increased risk of system failure. Speed increaser failure is typically caused by poor maintenance and insufficient lubrication (European Small Hydropower Association, 2004). Of all the types of speed increasers, the most efficient are generally high-grade roller chain and high strength synchronous belts (Francis, 2000). The advantage of the synchronous belt is the elimination of slip losses over a standard V-belt.

18

Figure 19: Speed Increasers Parallel shaft (top left), Bevel (top right), and Belt drive (bottom) (European Small Hydropower Association, 2004)

Turbine Erosion Consideration


The longevity and performance of a turbine is highly dependent on the erosion that occurs inside of the turbine. Such erosions that typically occur in a hydropower setting are due to cavitation, sand erosion, material defects, and fatigue (Kjlle, 2001). Furthermore, a form of erosion that is unique to the PRO process is metal erosion due the high concentration of salt that exists in the working fluid of the turbine. Filtration of both the fresh and saltwater inputs to the membrane housing will greatly reduce the concern for large amounts of sand erosion. Therefore, this section will primarily focus on cavitation and saltwater erosion inside the turbine. 19

Materials for Seawater Turbines Another important factor to consider for designing a PRO system turbine is the turbine material. The turbine must have adequate corrosion protection against seawater. Stainless steel alloys are typical choices for seawater pump application (Morrow, 2010), and it is assumed that application for turbines will be quite similar. The exact choice of alloy depends on several factors including the environment, seawater quality, desired lifetime of the pump/turbine, method of corrosion, whether it is the result of heightened velocity, galvanic effects, or biological damage (Morrow, 2010). The seawater quality parameters that must be considered are pH, salinity and chlorinity, temperature, conductivity, and quantity of dissolved gases and microbes. The most common alloy used in RO application, when the technology was still fairly new, was stainless 316L (Hassan & Malik, 1989). While not corrosion-proof in seawater application, it performs well under higher flow rates (British Stainless Steel Association, 2012). RO plants typically filter their water extremely well, causing the amount of bio-fouling and total suspended solids problems to be negligible, however, these must be considered for PRO application. Other stainless steels used in RO plants are 317L, 317N, 904L, and 254 SMO (Hassan & Malik, 1989). Behavior of certain alloys used in seawater pump applications is highly dependent on the velocity of the water. Type 316 (S31600) and nickel-copper alloys (N04400) perform the best for most applications. Unless pumps are left standing, crevice corrosion of 316 stainless is minimal (Todd, 1986). Nickel copper alloys, however, should not be used with steel or stainless steel cases because of the galvanic effects of the two metals (NiDI, 1995). A tin-bronze case is a better housing case alternative. A summary of the corrosion rates of various metals used for seawater pump application is shown in Table 1. The Copper Development Association reported very low biological fouling as another benefit of using a copper/nickel alloy for seawater piping, in this case it was a 90/10 ratio (Copper Development Association, 1986). Seawater desalination is one of the intended uses in the study. Drawbacks to using 90/10 copper-nickel alloy include lessened material strength, requiring thicker pipe walls for pressurized systems, and increased capital cost. Lifetime capital cost for copper-nickel is purported to be lower than that of high-grade stainless steel, however (Copper Development Association, 1986). The cost factors for various materials are displayed in Table 2 (Copper Development Association, 1986).

20

Table 1: Corrosion Rates of Common Materials Used for Seawater Pump Applications (NiDI, 1995)

Table 2: Relative Cost Factors of Various Materials for Seawater Piping Application (Copper Development Association, 1986)

Case Study: Materials Selection for a Japanese Pilot Seawater Pumped Storage Power Plant The materials selection for the turbine components of a pumped-seawater power plant on Okinawa Island of Japan depended primarily on the component and how it contacted the water. The plant utilizes a Francis-type pump-turbine and is designed for an operating head of 141 m and a flow of 26 m3/sec, producing a power output of 31.4 MW (Fujihara, Imano, & Oshima, 1998). Most of the wider, more easily accessible surfaces are made with rolled steel type SM400A, and this is coated with an extremely thick vinyl-ester paint layer with glass flaking (Fujihara, Imano, & Oshima, 1998). The components treated in this manner are the spiral case, parts of the head cover, and the general structure that does not contact water. The draft tube and parts of the head cover that had contact with water were made from type SUS316L austenitic stainless steel, and other components were made with SUS-F316N and 21

SCS16A steel (Fujihara, Imano, & Oshima, 1998). The entire assembly utilized an external power source for cathodic protection rather than a sacrificial anode. Material Hard Coatings for Resistance to Hydro-abrasion Hydro-abrasion is a function of many factors of the suspended solids in the water such as sand particle size, hardness, concentration, shape and velocity (Mann, 1999). Material hard coatings and surface treatments can be implemented to greatly increase the lifetime of a working component by increasing its resistance to abrasive wear. Several surface treatments have been empirically explored for use in regions in India with annual spikes in total suspended solids (Table 3).
Table 3 : Performance of Various Hard Coatings

T410 steel borided exhibited the greatest resistance to wear followed by D-gun sprayed 13Cr-4Ni WC + 12Co, followed by 13Cr-4Ni steel borided (Table 3). These surface treatments were tested under extremely high suspended solids loads that would typically only occur for several hours per year in a major storm event (Mann, 1999). Cavitation Cavitation is a phenomenon that occurs when the liquid pressure is less than the vapor pressure of a working fluid, causing vapor bubbles to form in the liquid. The vapor bubbles collapse as they move from a lower to a higher pressure zone, known as cavitation collapse, and can cause pitting and high levels of erosion inside the turbine (White, 2008). In the Francis turbine, areas most susceptible to cavitation erosion are the runners and draft tube cones. The degree to which cavitation will occur, in a given turbine, is found through a comparison of the Thoma coefficient, , and the Power Specific Speed, sp. As shown in Equation 33, the Thoma coefficient represents the ratio of the net positive suction head in a turbine to the overall available head (Dixon & Hall, 2010). The Thoma coefficient can also be described as the fraction of the net hydraulic head that is available for the production of work. ( )

Where HS is the net positive suction head in meters, HE is the total available hydraulic head in meters, Pa and Pv respectfully represent the atmospheric and vapor pressures in Pascals, is the density of the 22

(43)

fluid in kg/m2, g is the gravitational constant in m/s2, and z is the elevation difference between the turbine exit and the tail water elevation in meters. Inspection of the Thoma coefficient equation shows that at a given total amount of extractable head, it is necessary to optimize the net positive suction head, in an effort to minimize the amount of erosion due to cavitation. Special circumstances for the PRO project may leave the tail water elevation as the tidal elevation, which would result in a dynamic Thoma coefficient over time, and require a detailed analysis as to how this may affect the longevity of the turbine. The second value used to determine the likelihood of cavitation is the power specific speed, sp. The power specific speed for a given turbine can be calculated as follows: = ( )4
5

(44)

Where represents the power delivered to the shaft (Watts), and is the rotational speed in radians per second. Finally, with both sp and , the chart shown in Figure 20 is used to determine the likelihood of cavitation.

Figure 20: Cavitation versus specific speed used to determine degree of cavitation

Cavitation in Pelton wheel turbines is less of an issue, as the working fluid is at atmospheric pressure. However, cavitation in the Pelton wheel can occur in the nozzle and buckets of the turbine (Kjlle, 23

2001). Cavitation can also occur in impulse turbines at rough spots created by sand or silt erosion (Warnick, 1984). Given the exceptional quality of water supplied to the turbine, erosion due to cavitation on an impulse turbine in a PRO setting would likely be negligible.

Alternating Current (AC) and Power Factor (PF)


There exist numerous descriptions of AC theory and practical functionality. The reader may find more depth on this subject easily if desired. It is assumed for the rest of this discussion that the reader has a basic grasp of how voltage, current, and power behave in AC circuits and how they are referred to. An excellent reference for this material can be found at the Schneider Electric website in the design/application guide under electric power fundamentals. An important aspect of AC power systems is power factor. Power factor is the ratio of useful power expended at the load to the total amount of power delivered. The total amount of power delivered to the load will be larger than the real power consumed if the load includes reactive components such as inductors or capacitors. Inductors such as the coils in motors consume reactive power by phase shifting the current to a lagging position with respect to voltage while capacitors produce it by doing the opposite. Reactive power is important because despite causing no change in the real power consumed in the load it requires the presence of additional current to provide the magnetization necessary for inductive machines to work. This additional current causes transmission infrastructure to be loaded more heavily and incurs additional line heating losses. To compensate for the incurred costs from supplying reactive power many utilities levy fee penalties on power consumers whose load PF is below some threshold typically 0.8-0.85, this prompts the economic motivation for PF correction. Power factor correction uses capacitors or other electronic means to balance the reactive power consumption of inductors and reduce current flow. The capacitance required to offset a given inductance is simple to calculate but the practice of power factor correction is so widespread that is common for units to simply specify the amount of reactive power they can provide in kVAR (kilo-volt amps reactive). Commercially available PF correction banks often include electronic control circuits that include necessary protection systems. Correction banks are usually placed in parallel with the individual load components terminals (static PF correction) but can also be placed in bulk in parallel with the entire load. In bulk PF correction PF monitoring equipment is constantly switching capacitor banks to maintain the corrected power factor above a set point often 0.95 and no hazards associated with accidental unity PF or leading PF (larger capacitive loading). In static PF correction, no monitoring equipment is required as the PF correction bank is connected in parallel with the motor and gets switched on when the motor does. It is critical however that the correction not bring the power factor to unity or to a leading state do the potential for damagingly high resonant current to develop when the motor is disconnected from the line (Photonics LTD., 2002).

24

Electro Mechanical Equipment


To deliver electrical power to a load the power in a turbine shaft must be converted to electricity with a generator and conditioned for use and safety with switchgear, transformers, and breakers. Generators The simple generator takes advantage of a changing magnetic flux through a coil to induce current in a wire coil. This change in flux caused by rotational motion of a rotor within a fixed structure called a stator. There are many configurations of generators that specify if the magnetic field is generated by a permanent magnet or an electromagnet. Configurations change depending on the location of the magnet (on stator or rotor), the number of magnetic dipoles, and the particular schematics for the windings. For the purpose of this project it is important to discuss the configuration and operation of AC generators that can be generally divided between synchronous and asynchronous types (Schultz, 1985). Generators can theoretically be manufactured to produce any number of distinct phases, but in practice, either single or three phase generators are produced. The primary reason to select a three-phase generator over a single phase one is if the load requires power in three phases as in the case of large motors (Norwol Power Systems, 2012). Synchronous Generators Synchronous Generators use a rotor mounted electro magnet to induce current in the stator windings (Figure 21). This configuration reduces the difficulties associated with mechanical operation of the rotor due by decreasing rotor weight. The configuration also simplifies the requirements for current carriers between the rotor and the rest of the system. The exciter supplies the current required for operating the electro magnet. This current is typically much smaller than the load current, requires a smaller conduit, and reduces the likelihood of resistance heating failure (Raabe, 1985). The exciter can be powered by the generator itself (self-excitation) or by an external power source through brushes that contact the rotor. When a generator is set up for self-excitation brushes are still included to contact the rotor so the generator can be excited externally in case of a failure (Raabe, 1985).

25

Figure 21: Three Phase Synchronous Generator (PDH, 2012)

Rotor Speed in Synchronous Generators Rotor speed is an important performance characteristic of synchronous generators. A rotor can have either one or multiple magnetic dipoles on it and the synchronous speed of the generator can be manipulated by selecting the number of poles. The mechanical speed of the rotor is related to the number of poles on the rotor and the frequency of the output (Eq. 35) (Schultz, 1985). = 120 (45)

Where is the mechanical speed of the rotor (rpm), is the output power frequency (Hz), and is the number of poles per winding (for a single dipole = 2). The relationship above shows that the construction of a generator that can operate at lower rotational speed is possible. This mode of operation is suitable for hydroelectric application (Schultz, 1985).

26

Asynchronous (Induction) Generators Induction generators are simpler and cheaper to construct in comparison to synchronous machines. The excitation for induction generators is provided by the AC signal on the grid. This means that induction motors consume reactive power because the magnetizing excitation current is provided by the grid. Often capacitors are installed parallel with them to provide the reactive magnetizing current as power factor correction for the generator. Induction generators require less maintenance than synchronous ones and tend to require less complex electronic controls. Induction generator terminal voltage and frequency is held constant by the grid (Heng, 1992).

Figure 22: Torque and Speed characteristics of an induction machine (Heng, 1992).

When an induction machine is driven above synchronous speed it acts as a generator delivering power to the grid (Figure 22). The machine must be driven above synchronous speed by a prime mover. The relationship between the rotor speed in an induction machine and the synchronous speed for the grid is called the percent slip. Operating power is typically achieved at approximately 3% slip. Induction machines are particularly vulnerable to over-speed conditions so careful attention must be paid to the speed governor in the induction generator set (Shahl, 2013).

Generator Selection The optimal generator selection can be deduced from assessment of several design parameters. Initially examination of speed and power requirements may be enough to inform generator selection (ABB, 2011) (Figure 23). If the parameters fall in the overlapping section then a more careful assessment is required (ABB, 2011). The following feasibility studies should be considered to select an optimal generator: An economic analysis should be conducted to determine if added efficiency from a synchronous machine will pay for its additional capital investment over the project lifetime. On site requirement for power factor correction should be examined. The capital investment required for power factor correction should be included in the economic analysis.

27

Figure 23: Selection of a Generator Informed by the Operating Speed (of which the number of poles is a function) and the Power Output Available from the Prime Mover (ABB, 2011).

Governors Governors monitor the load conditions of the system and makes adjustments to the system to maintain stability and functionality. These adjustments can be made either by changing the hydraulic characteristics upstream of or through the turbine to reduce the hydraulic power delivered or by delivering excess power to shunt loads. Very advanced digital governing systems are available that constantly monitor load characteristics, rotor speed, cooling systems, and other system parameters and adjust everything automatically. These systems also typically include all switchgear necessary for the generator (HAP, 2011). Transformers Transformer theory is well reproduced in many locations, and an excellent source for this material can be found in the Hydropower Advancement Projects best practices catalogue under transformers. In small-scale power generation facilities, a transformer will usually be required to step down the voltage produced by the generator to the onsite line voltage. Typical small (size relevant to project) generators in the U.S. operate at 480 kV or 690 kV. Line voltages for most end use loads are 120 V or 240 V. Switchgear Switchgear is a blanket term for all of the protective equipment that breaks the circuit if something goes wrong. Circuit breakers provide overcurrent protection for components such that high current from a fault or over speed does not cause resistance heating that damages the equipment. It is important that breakers be properly sized so that they provide adequate protection without tripping unnecessarily. When generators are designed to run at their rated load constantly, it is necessary to employ a digital 28

circuit that monitors temperature in the generator coils and trips a breaker if a temperature threshold is exceeded.

Power Quality and Conditioning


Successful and efficient grid network operation requires that the voltage and frequency supplied meet the demands of the load. Generators have the ability to affect these properties of the load more as they supply a greater proportion of the total load. In the case of distributed generation the load is the grid and the generator tends to be a small portion of the total generation. The main power quality problems commonly encountered in practice are listed and defined by the Institute of Electrical and Electronics Engineers (IEEE) as the following (Brown, 2006): Overvoltage: An RMS increase in the AC voltage, at the power frequency, for a period of time greater than 1 min. Typical values are 110%-120% of nominal. Undervoltage: An RMS decrease in the AC voltage, at the power frequency, for a period of time greater than 1 min. Typical values are 80-90% of nominal. Swell: An increase in RMS voltage or current at the power frequency for durations from .5 cycle-1 min. Typical values are 110%-180% of nominal. Sag: An RMS reduction in the AC voltage, at the power frequency, for durations from cycle to a few seconds. Interruption: The complete loss of voltage. A momentary interruption is a voltage loss (<10% of nominal) for a time period between .5 cycles and 3 seconds). A temporary interruption is a voltage loss (<10% of nominal) for a time period between 3 seconds and 1 min. A sustained interruption is the complete loss of voltage for a time period greater than 1 min. Notch: A switching (or other) disturbance of the normal power system voltage waveform, lasting less than _ cycle; which is initially of opposite polarity to the waveform, and is thus subtractive from the normal waveform in terms of the peak value of the disturbance voltage. This includes a complete loss of voltage for up to _ cycle. Transient: A subcycle disturbance in the AC waveform that is evidenced by a sharp discontinuity of the waveform. May be of either polarity and may be additive to, or subtractive from, the nominal waveform. Flicker: A variation in input voltage, either magnitude or frequency, sufficient in duration to allow visual observation of a change in electric light source intensity. Harmonic Distortion: The mathematical representation of distortion of the pure sine waveform. This refers to the distortion of the voltage and/or current waveform, due to the flow of non-sinusoidal currents. Electrical Noise: Unwanted electrical signals that produce undesirable effects in the circuits of the control systems in which they occur. Noise may be further categorized as transverse-mode noise, which is measurable between phase conductors but not phase-to-ground, and commonmode noise, which is measurable phase-to ground but not between phase conductors. This noise may be conducted or radiated. Also referred to as RFI (radio-frequency interference) or EMI (electro-magnetic interference). 29

Table 4: Common causes of power quality problems (Brown, 2006)

Operation and Maintenance


Since there are no currently operating PRO facilities of similar size to this project. Operation and maintenance (O&M) costs for the powerhouse of small hydroelectric generation facility (< 1MW) were used to derive the O&Mcosts of this projects powerhouse. The powerhouse includes the turbine, the generator and the electrical protection and switchgear. Hydroelectric generation facilities of similar size to this project (<1MW) have an annual operation and maintenance cost of 1.5-2.5% of the initial capital investment for the project (Lako, 2010).

30

Regulatory Framework
This project is subject to federal, state, and local regulations in addition to remain in compliance with the California Environmental Quality Act (CEQA) process that will involve the collaboration of numerous state and federal agencies.

Federal Regulatory Agencies


The project is subject to the Federal Energy Regulatory Commission (FERC) guidelines for energy generation facilities. FERC regulations require new energy generation facilities to complete a license application. The project may also be subject to the Bureau of Ocean Energy Management (BOEM) since the seawater is used in the PRO process (Murray, 2013). Beyond providing licenses to operate the energy generation facility, these federal agencies may become lead agencies during the CEQA process.

State Regulatory Agencies


The project is subject to the California Energy Commission (CEC) regulations and must cooperate with the California Public Utility Commission (CPUC) during the CEQA process. Relevant State Laws and Decisions California Renewable Energy Resources Act (Senate Bill X1-2) Senate Bill X1-2 from 2011 established the California Renewable Energy Resources Act, setting the goal of 33% renewable energy generation by 2020. The CEC is responsible for reviewing renewable energy applications. Though PRO has been scientifically established as a renewable energy source, there are no existing facilities in the United States, and the PRO project may have to provide additional information beyond the renewable energy application to prove that it meets the guidelines of a renewable energy generation facility under current legislation. Renewable energy generation facilities are eligible for incentive programs like the Emerging Renewables Rebate Program that offers cash rebates to qualified generators (EPA, 2013). Rebates offered by this program will help to reduce the payback period of the project and help bring the project towards economic feasibility. CPUC Decision (D.)11-07-031 The CPUC has allowed an extension of virtual net metering to all multi-tenant properties, which allows the bill credits associated with the electricity produced by the system to be distributed to all the tenants electricity bills (CPUC, 2011). This decision will allow the Humboldt Bay Harbor Recreation and Conservation District (HBHRCD) to sell power to leases of the industrial park instead of selling the power to PG&E. This would allow the HBHRCD to sell power to their tenants at slightly less than market value instead of entering into a power purchase agreement with PG&E where they would sell the power for much less. 2011 CPUC Market Price Referent (MPR) The CPUC Market Price Referent (MPR) represents a levelized energy price where expected generation revenues equal expected generation costs with respect to a net present value analysis. The MPR was established by the CPUC through a series of decisions based on Senate Bill 1078 and Senate Bill 107. A 31

model was developed using Combined Cycle Gas Turbine energy generation facilities as the proxy. The model allows the input of a projects start year and expected lifetime and will produce estimate energy price values for each year of the life cycle to aid in project planning. Using the project start year of 2013 and a 20 year life cycle, the first year energy price was predicted to be $0.09375/kWh.

Local Regulatory Agency


Pacific Gas and Electric (PG&E) is the local utility that can provide various grid connection scenarios based on state regulations defined by the CEC and the CPUC. 2013 Renewable Market Adjusting Tariff (ReMAT) Power Purchasing Agreement The CPUC Decisions (D.) 12-05-03 and (D.) 13-05-034 established PG&Es Electric-Renewable Market Adjusting Tariff (E-ReMAT) on July 24, 2013 to implement Senate Bill 32. Senate Bill 32 increased the statewide renewable procurement target from 500 MW to 750 MW and increased eligible project size from 1.5 MW to 3 MW (AC). The contract price for ReMAT eligible Power Purchasing Agreements will begin at $89.23/MWh. (PG&E, 2013)

Grid Connection and Virtual Net Metering


PG&E provides a virtual net metering rate schedule (NEMV) for multi-tenant or multi-meter properties that have a renewable energy generation facility with a single service delivery point (SDP) connected to PG&Es utility grid. A renewable energy generation facility is defined by the California Energy Commissions (CEC) Renewables Portfolio Standards (RPS) guidebook to be a generation facility that generates electricity by using the following certified generation technologies: biomass, solar thermal, photovoltaic, wind, geothermal, fuel cells using renewable fuels, small hydroelectric generation, digester gas, municipal solid waste conversion, landfill gas, ocean wave, ocean thermal, or tidal current. Pressure Retarded Osmosis (PRO) has not yet been established as a renewable energy technology in the RPS guidebook. A CEC-RPS-1B application will need to be submitted to the CEC to apply for precertification during the planning phase of the project (CEC, 2008). It is very likely our project will be certified as a renewable energy generation facility based on personal correspondence with the CEC (Loyer, 2013). Once PRO has been certified as a renewable energy generation technology under the RPS, the ability to operate under PG&Es NEMV rate schedule will be possible.

32

Power Sales
The Humboldt Harbor District can choose from two scenarios to sell the power generated on site in the PRO facility: Utilize virtual net energy metering to integrate energy usage costs of future industrial tenants into lease agreements as a surcharge Sell the power to PG&E under a ReMAT power purchasing agreement

Operating under the guidelines of PG&Es NEMV rate schedule, the Harbor District could set up multiple PG&E accounts (one for each industrial tenant) in addition to a main account connected to a single SDP. A PG&E smart meter will be associated with each account (for each tenant) allowing the Harbor District sell power produced on site to the industrial tenants as part of their lease agreement for slightly less than market value. This would be an incentive to potential tenants to be able to purchase a portion or possibly all of their power at below market value. One scenario would break up the lease into a cost for the building space, and energy usage costs. The first several months would serve as an assessment period where energy use could be determined and applicable charges applied to the lease agreement. The language of the lease agreement should allow for periodic reassessment of energy usage costs to allow for operational changes or efficiency improvements of the tenant. Operating under a ReMAT power purchasing agreement (PPA), the Harbor district would be able to sell power for $0.089/kWh beginning in November of 2013. Eligibility for a ReMAT (PPA) is subject to the approval of the project application under the Renewables Portfolio Standards (CEC, 2008). Additional analysis of the time of use schedule should be conducted such that load schedules which take advantage of the cheapest rates are implemented. Due to the non-availability of information concerning the cost of fixing and maintaining the existing transformer station on site another important analysis was not possible but merits further attention. The existing transformer connects to the grid at very high voltage which allows the site power consumption to be billed as a primary consumer. Primary consumption purchase rates are lower than those for secondary consumption. On site consumers would likely be more prepared to utilize power from a secondary voltage level source. This presents the potential opportunity for the site to operate the existing transformer as a revenue source regardless of on sight production. Analysis would be needed to determine if the cost of repairing and maintaining the transformer would cost less than the available revenue from slight markup of the power sold to tenant consumers (from primary rates to secondary rates).

33

Turbine Component Design


Constraints and Assumptions
In coordination with the other project groups, the optimum design pressure and design flow are about 1550 kPa and 0.50 m3/s, respectively, though some small variation is expected. In addition the plant is assumed to run at peak flow continuously for analysis of its performance. The current estimate of the system parasitic load range is based on water pretreatment options, and found to be between 305 kW and 598 kW. The system was initially expected to require some amount of residual head to aid in pumping effluent through the outfall pipe itself, but calculations performed by the group designing the outfall show that pumping the outfall will not be necessary, and gravity feed from the wet well at the northwestern corner of the property is sufficient for a 30 MGD flow. However, pumping is required to deliver diluted seawater from the turbine outfall to the wet well, so residual head may indeed be an important consideration. The site has an estimated ground elevation change of 13 feet between the location of the turbine and the wet well, using Google Earth path and elevation profile tools. The following assumptions were made for the economic analysis portion: A discount rate of 4.0% over the life of the project The project life is 20 years An annual rate increase that follows predictions made by the CPUC MPR model Additionally, all costs were brought to present value in 2013 dollars for the easiest economic comparison between alternatives. A summary of the constraints and assumptions described is presented in Table 5.
Table 5: Design constraints for selection of the hydroturbine.

Assumption/Constraint Design Flow Design Pressure Pumping to Wet Well

Value 0.5 m3/s 1550 kPa 2 m head

Detail Dictated by output from membrane optimization model Dictated by output from membrane optimization model Assumed necessary to overcome elevation difference

The maintenance schedule for the PRO membranes is expected to keep the plant functioning at the design pressure and flow. Redundancy in the number of membrane modules will provide opportunity for servicing while the plant is fully functional. The plant is also assumed to operate 365 days of the year.

34

Selection Criteria
The important criteria to consider when matching a hydro-turbine to the PRO plant, in order of importance, are: Sufficient power output so that net production is positive (technically feasible) Overall efficiency Lifetime net cost (including capital cost, operating cost, and revenue) Ability to retain head in turbine outflow risk of cavitation damage Range of efficient operation for part-flow

The first criteria must be met for any alternative to be considered. The efficiency, costs, and revenue of each system will then be estimated and used to determine the most cost-effective option. The ability of the turbine to retain residual head is important for pumping the outfall, because it may provide the most cost-effective pumping option. The final criterion calls for assessment of the part-flow efficiency of each turbine. As described in the literature review, the impulse turbines selected as alternatives generally have greater part-flow efficiency than the reaction alternative. Based on the redundancy and servicing schedule developed by the membrane design team, however, this criterion is a very minor consideration. The selection criteria, including applicability of the turbine design, efficiency concerns, and capital cost, are tabulated for each alternative near the end of the section (Table 16).

Turbine Design Alternatives


Based on typical operating ranges from the literature, four turbine types are applicable for the facility. These are the Pelton Wheel, Turgo Runner, and Cross-flow impulse turbines, and the Francis reaction turbine. The designs have been weighed against the criteria examination, and preliminary designs for the Pelton and Francis were developed. Other turbine types, including the Kaplan and fixed-vane Propeller, were not considered because the design head and flow were well outside their typical operating range. The turbine alternatives were researched and preliminary designs were developed for each. The performance of these designs was compared to determine the preferred alternative. The economics for each alternative were developed assuming that all the generated power was sold to the utility at the Annual MPR produced by the CPUC model. The average annual rates were multiplied by the net annual energy output to compute an optimistic (plant operates 24 hours and 365 days per year) amount of revenue that could be obtained by selling the power. The net energy was calculated by subtracting parasitic loads in each section of the plant on an annual basis. This methodology is rough and it is certain that intentionally operating periodic loads during off and partial peak hours would improve the economics. Under the limitations of this design schedule however, coordination of an in depth hour by hour analysis for power consumption schedule optimization across 5 subsystems was unfeasible. The E-19 schedule was determined to produce a more favorable (higher) time averaged rate. Pelton Wheel Design The preliminary Pelton turbine design developed for the project has an expected overall efficiency of approximately 84% and may produce 695 kW of electrical power. The design was developed from the kinetic theory described by the literature. The system assumptions are displayed in Table 6..

35

Table 6: Pelton Turbine Design Parameters

Parameter Nozzle Losscoef. (n) Bucket Discharge Angle () Bucket Loss coef. (m) Desired Rotational Speed (rpm) Generator Efficiency Transmission Efficiency

Value 0.97 165 0.95 600 95% 97.5%

Reasoning Typical of well-designed systems (Warnick, 1984) Typical value (Warnick, 1984) Typical value (Warnick, 1984) Typical (Harvey, 1993) Typical for size High, extremely well designed (Francis, 2000)

The generator efficiency for the system was conservatively assumed to be 95% for an induction generator operating as part of a system of this size. The calculations for the turbine design values are shown in the Appendix along with the Excel sheet used. The calculations produced an optimal Pelton wheel design with the characteristics shown in Table 7.

Table 7: Pelton Design Outputs

Design Output Output Power (shaft) Turbine Efficiency Specific Speed Number of Jets Jet Velocity Runner Velocity Runner Diameter Number of Runner Buckets Bucket width Bucket Radial Length Bucket Depth Bucket Spacing Torque Specific Energy Extracted Overall Efficiency Generator Output Power

Value 695 kW 90.1% 0.189 2 52.7 m/s 25.2 m/s 0.80 m 21 0.20 m 0.17 m 0.07 m 0.12 m 11.1 kN-m 0.376 kWh/m3 83.8% 644 kW

The expected generator rotational speed for the Pelton design is between 600 and 1200 rpm. To directly couple the devices, the generator would have to run at 600 rpm, same as the turbine. The number of poles required for a generator running at 600 rpm is 12, which is uncommon, and it is below the standard speed of typical alternators (European Small Hydropower Association, 2004). A speed increaser could be used to bring the generator rotational speed to 1200 rpm, but this would sacrifice a small amount of efficiency. A synchronous belt speed increaser is the best option for application with a Pelton turbine of this size. 36

Advantages, Disadvantages, and Cost of the Pelton Wheel The advantages and disadvantages of the Pelton wheel alternative are displayed in Table 8.
Table 8: Advantages and Disadvantages of the Pelton Wheel

Advantages Good part-flow efficiency Unpressurized housing for turbine Simpler construction Lower risk of cavitation

Disadvantages No possibility for residual head Lower maximum efficiency

The primary advantages are the fact that it has the best part-flow efficiency of any turbine. If the plant runs at a flow that is less than the design flow, a Pelton turbine will still produce power nearly as efficiently as it would at full design flow. The efficiency declines rapidly after a threshold of approximately 25% of full design flow. Because a Pelton wheel does not require a pressurized housing, the construction cost is lower and maintenance is simpler. Cavitation is only a problem for Pelton turbines when silt or sand in the water begins to pit the runner, allowing small areas for potential cavitation damage. Cavitation damage of this sort is expected to be extremely minimal considering the filtration requirements of the incoming water, so this is an advantage of the Pelton design. A capital cost estimate for a Pelton turbine for the PRO system was provided by Canyon Hydro, totaling between $500,000 and $600,000 for the turbine, generator, and power conditioning equipment. This price includes the engineering and installation, but it does not include the powerhouse building. The company representative also relayed that the designed turbine would run at 600 rpm, which is consistent with the calculated preliminary design (Canyon Hydro Representative, 2013). A second estimate using a very recent cost curve from a European study was calculated at $568,000 (SWECO Norge AS, 2012). Turgo Runner Design The preliminary design for the Turgo Runner has an expected overall efficiency of 81% and 655 kW of power output from the generator. The design was developed from kinetic theory similar to the Pelton design. The system inputs are displayed in Table 9.
Table 9: Turgo Turbine Design Parameters

Parameter Nozzle Loss coef. (n) Bucket Entry Angle (1) Bucket Exit Angle (2) Bucket Loss coef. (m) Desired Rotational Speed (rpm) Generator Efficiency

Value 0.97 20 10 0.95 1200 95%

Reasoning Typical of well-designed systems (Warnick, 1984) Typical (Crewdsonem, 2013) Typical (Crewdsonem, 2013) Typical value (Warnick, 1984) Typical, twice similar Pelton (Nasir, 2013) Typical for size

37

The generator efficiency assumed is the same as for the Pelton turbine. Because the specific speed of the Turgo turbine is higher, it is possible to directly couple the turbine to the generator, eliminating the need for a speed increaser and improving efficiency. The Turgo wheel designed would have to be coupled to a six pole generator and run at 1200 rpm. The outputs from the design calculations are shown in Table 9.
Table 10: Turgo Design Outputs

Design Output Output Power (shaft) Turbine Efficiency Specific Speed Number of Jets Jet Velocity Runner Velocity Runner Diameter Number of Runner Buckets Bucket Spacing Torque Specific Energy Extracted Overall Efficiency Generator Output Power

Value 655 kW 84.9% 0.378 2 52.7 m/s 23.7 m/s 0.38 m 20 0.059 m 5.22kN-m 0.354 kWh/m3 80.7% 623 kW

Compared to the Pelton Wheel, the Turgo Runner design has a lower efficiency. It does have similar advantages of decent part-flow efficiency and lower risk of cavitation damage. It also does not require a pressurized housing, but it also cannot provide residual head. No reliable source was found that allowed a confident cost estimate for a Turgo turbine, so it was assumed that a Turgo runner design will be nearly equal to the cost of a Pelton Wheel because of the similarities in their design. Advantages, Disadvantages, and Cost of the Turgo Runner The advantages and disadvantages of the Turgo Runner are displayed in Table 10.
Table 11 : The advantages and disadvantages of the Turgo Runner

Advantages Good part-flow efficiency Unpressurized housing for turbine Lower risk of cavitation

Disadvantages No possibility for residual head Lower maximum efficiency Limited manufactures

Crossflow Turbine Design The crossflow turbine design is estimated to have an overall efficiency of 80% with an expected output of 648 kW. The design was developed using a spreadsheet tool provided by a micro-hydro installation, and also incorporated additional parameters based on experiments performed by Professor Nadim Aziz of Clemson University. The input parameters are given in Table 12.

38

Table 12: Crossflow Turbine Design Inputs

Parameter Exp. Engineered Turbine Eff. Runner Outer Diameter (in) Diameter Ratio (D2/D1) Angle of Attack Blade Exit Angle Speed Ratio Number of Blades Desired Rotational Speed (rpm)

Value 84% 23.5 0.67 22 105 0.7 20 1200

Reasoning Typical for engr. systems (Aziz & Desai, 1991) Allowed direct coupling, smaller than typical for size 0.6 0.7 is best (Aziz & Desai, 1991) Typical >90 (90 is ideal) (Zia, Ghani, Wasif, & Hamid, 2010) 0.54-0.67 typical ideal range (Aziz & Desai, 1991) Low end of range (Aziz & Desai, 1991) Allows direct coupling of generator if rpm of turbine

The crossflow design model was not a sophisticated as the others and did not provide an estimated efficiency as a function of the inputs. Efficiency had to be assumed based on the literature. The design basically calculated the sizing of the turbine runner, blades, and some of the angles. A fairly generous expected efficiency was assumed, but the overall design still fell behind the other two. The design outputs are given in Table 13. The values have been converted to metric units from the Imperial units of the calculator tool.
Table 13: Crossflow Turbine Design Outputs

Design Output Output Power (shaft) Turbine Efficiency Jet Velocity Runner Velocity (at edge) Metric Specific Speed Runner Outer Diameter Runner Inner Diameter Blade Spacing Gear/Belt Ratio Specific Energy Extracted Overall Efficiency Generator Output Power

Value 648 kW 84% 54.3 m/s 37.5 m/s 57.9 0.37 m 0.60 m 0.094 m 1.00 (allows direct drive) 0.350 kWh/m3 79.8% 615 kW

The crossflow turbine design used a relatively small runner diameter for the size flow to allow direct coupling of the generator as a design option. This would eliminate transmission drive losses. There was some skepticism whether or not the turbine could operate efficiently at 1200 rpm, but the calculated metric specific speed was still well within the desirable range for a cross-flow turbine (Mathema, 2013). The cost of crossflow turbine for the system was estimated at $575,000 using a relatively old study on turbine costs (U.S. Army Corps of Engineers, 1979). The total cost includes the turbine, generator, speed governing equipment, and installation. Advances have been made since the 1970s in cross-flow turbine design, and they are often fabricated quite cheaply in third-world countries, so the cost may be substantially lower than this estimate. Still, it falls in the expected range for a system of this size. 39

The Advantages and disadvantages of the cross flow turbine The advantages and disadvantages of the cross flow turbine are displayed in Table 13.
Table 14: Advantages and disadvantages of the cross flow turbine

Advantages Good part-flow efficiency Unpressurized housing for turbine Low risk of cavitation

Disadvantages No possibility for residual head Lower maximum efficiency Limited manufactures

Francis Turbine Design A preliminary Francis Turbine design was provided by software output from Canyon Hydro. Canyon Hydro custom designs their turbines to the head and flow of the site and specializes in small hydro systems. Using the site parameters, the representative provided the following design specifications for a Francis Turbine for the site (Table 15).

Table 15: Francis Turbine Design Specifications

Design Output Turbine Efficiency Generator RPM Overall Efficiency Electric Power Output

Value >90% 1200 87% 671 kW

Canyon Hydro also provided a cost range for the entire system, including the turbine, generator, power conditioning equipment, and installation, totaling $800,000 to $1,200,000 (Canyon Hydro Representative, 2013). The operational rotational speed of the turbine was also determined to be 1200 rpm, which makes direct coupling of the turbine and generator possible and therefore more efficient and cost-effective. The turbine design also allowed for swirl in the draft tube of the turbine which produced residual head for the outflow (Canyon Hydro Representative, 2013). A design that did not induce swirl has an expected efficiency of over 90%.

40

Preferred Alternative
Comparison of Alternatives with Selection Criteria The four alternative designs are tabulated for comparison in Table 16 with the factors that relate to each of the selection criteria. After just preliminary analysis, the Pelton impulse turbine alternative is shown to outperform both the Turgo and Crossflow impulse turbines.

Table 16: Summary Table of Alternatives and Selection Criteria

Turbine Type Turgo Crossflow Pelton Francis

Flow Range (m3/s) 0.05 -5 0.1 - 13 0.02 - 12 0.1 - 30

Head Range (m) 15 - 300 15 - 300 30 - 1300 3 - 300

Overall Efficiency (%) 81 80 84 83

Capital Cost $550,000 $575,000 $568,000 $1,000,000

Lifetime Revenue $864,134 $784,593 $1,112,724 $1,426,642

Part Flow Residual Efficiency Head Good Good Great Marginal No No No Yes

It is necessary to move the water exiting the turbine uphill to the outfall wet well. The only alternative that fulfills this requirement passively (without additional pumping) is the Francis turbine. The Pelton impulse turbine is expected to be more efficient, however, so a cost analysis was completed that compared the capital and energy cost savings of using the Francis turbine over the Pelton turbine (Table 17). The other impulse turbines were not considered because the Pelton wheel outperforms both of them for almost every criterion.
Table 17: Comparison of Pumping Cost using Pelton Wheel vs. Residual Head using Francis Turbine

Turbine Pelton Francis

Overall Efficiency (%) 84 83

Capital Cost $568,000 $1,000,000

Lifetime Revenue $1,112,724 $1,426,642

Lifetime Pumping Cost $302,588 0

Net over Lifetime (Revenue Cost) $242,136 $426,642

Pump capital costs for this option were priced at $50,000 using several smaller pumps due to the high flow and low head required. The energy cost was also calculated over the life of the pumps. These extra costs associated with the Pelton wheel (and any other impulse turbine options) make the lifetime value of the Pelton turbine lower than that of the Francis turbine. This demonstrates that the Francis turbine is the better option despite its much higher capital cost high capital cost. Choice of Preferred Alternative The Francis Turbine was chosen as the preferred alternative based on the following qualifications: Best lifetime economics of any alternative Highest efficiency 41

Highest lifetime net revenue Eliminates the need for an extra pump to move water to wet well

Although the lifetime economic analysis conducted did not take into consideration the rest of the plant, all the costs described in the other sections were assumed not to change based on the turbine and generator selection. Because the optimum design flow and head are determined by the membrane group, and the other groups used these parameters to size their systems as well, the turbine selection is independent of the cost of the other systems. Selection of System Generator The selection of the generator is the primary choice made about the electromechanical configuration of the system. The main considerations are the generator type (synchronous vs. asynchronous), number of poles, size match with the system, rated power and rated voltage (Table 18).
Table 18: Generator Specifications

Generator Type Induction

# of poles 6

Rated Power 700 kW

Rated Voltage 440

The reason for selecting an induction generator is motivated by the reduction in capital cost and maintenance cost associated with this option. It is most effective to rationalize this design decision in the context of operational considerations that might prompt the adoption of a synchronous generator. Synchronous generators provide the operator with the ability to operate disconnected from a central power grid; they offer the capability to provide grid voltage regulation and deliver reactive power to magnetize inductive loads. The project is grid connected and small enough with respect to local power demand that we would make an insignificant contribution to grid voltage regulation. There are inductive loads on site but due to the nature of the production schedule it would be undesirable to produce the reactive power with a synchronous generator. The production schedule is a constant base load power output which is limited by the hydraulic power available. Reactive power is produced in a synchronous generator by overexciting the rotor coils which causes a leading power factor in the generator. When operating at the generators rated load the limiting parameter is not the real power out but the current in the coils so providing reactive power limits the real power output from the machine. The reactive power required on site to prevent power factor charges on the power bill can be provided with static power factor correction banks. Additionally the project is not large enough in scale that a synchronous generator would be more efficient. Due to these factors there is no motivation to make a larger investment in capital and O&M costs for a synchronous generator.

42

Design Detail
The summary design of the preferred Francis turbine alternative is presented with an image of a similar system, physical description of the key required components, and a schematic of the piping conveyance to the wet well. Installing the turbine with a horizontal axis of rotation would be preferred because of the geometry of the facility piping (Figure 24). The main collection pipe is designed to be routed into the powerhouse area near ground level to minimize elevation head losses, and this configuration also allows the draft tube to be routed to minimize additional head loss after the turbine. This is critical because of the head required to move the water to the wet well.

Figure 24: Example system with selected design characteristics (Hydrolink, 2009)

A design for the pip routing between the turbine and the wet was developed (Figure 25). The total elevation head between the turbine draft tube and the wet well is 13 feet (about 4 meters), and it was assumed that a total of 2 meters of head must be provided by the design of the Francis draft tube to pump flow to the wet well. This is assuming that the wet well outlet is at least 2.5 meters deep. The head loss between the turbine and the wet well is designed to be less than 0.5 meter. The short length of pipe between the turbine/ERD and the wet well is sized at 24 leaving the turbine draft tube and 36 after the junction with the ERD piping. The pipe material is low grade stainless steel, guaranteed for a lifetime of 20 years. The pricing for the pipe system was calculated using a tool provided by Taylor Engineering (Taylor Engineering, 2013).

43

Figure 25: Proposed general site layout

A summary of the final design specifications for the system are shown in Table 19. A simplified schematic of the horizontal layout of the turbine and generator itself is shown in Figure 26.

Table 19: Parameters of final design

Design Component Francis Turbine RPM Output Generator Power Equip. Pipe to Wet Well Material Length

Size/Type 671 kW 1200 671 kW 700 kW 36 Steel 1485 feet

Cost $1,000,000

Source (if applicable) Custom designed by Canyon Hydro

Included in system Included in system $1,070,000

Part of package (Induction gen.) Part of package Taylor Engineering Pipe Sizing Tool Life cycle of 20 years Includes contractor markup and labor

44

Figure 26: Simple diagram of the horizontal Francis turbine and generator setup

The total capital cost for the turbine and generator system, including piping to the wet well, is $2,070,000. The annual O&M is expected to be $44,000. The expected revenue for the alternative is between $189,000 and $235,000 annually, which resulted in a total net present value of roughly $820,000 for the turbine and generator system.

45

Economics of the PRO Power Plant


The economic analysis for the entire plant was conducted in terms of net present worth. Two separate analyses were conducted to account for the differences between the freshwater 1 and freshwater 2 teams designs. An aggregate capital cost for the project was calculated by adding the capital costs of each subsystem (Figure 27). An aggregate O&M cost was calculated by adding the O&M budgets for each subsystem, this O&M was assumed to remain constant over the 20 year life of the plant (Figure 28). Each year of O&M was projected into present worth using a discount rate of 4% (agreed upon between groups for consistency), and the sum of each years present worth was calculated as the present value of the O&M costs for the plant.

Figure 27: Aggregate capital costs for the PRO power plant

46

Figure 28: Aggregate O&M costs for the PRO power plant

Computation of the revenue required an analysis of the net power available from the plant. This entailed determining the energy consumed by each plant sub-system to determine the parasitic loading of the plant (Figure 29). Subtracting the parasitic loads from the gross power production yields the net sellable power. This power was assumed to be output constantly for one year to determine the sellable net energy and compute the annual revenue.

Figure 29: Net power determination is based on parasitic loading in each plant sub-system; two analyses show the variations in design conclusions from the freshwater 1 and freshwater 2 groups.

47

It was assumed that net energy could be sold at the rates specified by CPUCs MPR model so revenue in each year was computed as the product of the net energy and the predicted rate for the particular year. Each years revenue was brought back to present worth with a discount rate of 4%. The net present value of the entire project was then computed by subtracting the net present value of aggregate O&M costs and the aggregate capital costs from the net present value of the revenue. The net present value of the plant is between -$62 million and -$64 million. For more significant figures and example calculations see the Appendix.

Conclusions
The Francis Turbine was chosen as the preferred alternative based on the following qualifications: Best lifetime economics of any alternative Highest efficiency Highest lifetime net revenue Eliminates the need for an extra pump to move water to wet well

Paired with the Francis turbine is a three-phase induction generator. The reason for selecting an induction generator is motivated by the low capital cost and O&M cost associated with this option coupled with the lack of operational requirements that might justify upgrading to a synchronous machine. The total capital cost for the turbine and generator system, including piping to the wet well, is $2,070,000. The annual O&M is expected to be $44,000. The expected revenue for the alternative is between $189,000 and $235,000 annually, which resulted in a total net present value of roughly $820,000 for the turbine and generator system. High on site energy consumption for moving and treating water are the largest contributors to making reducing the net energy output and making this project infeasible economically. Future analysis of PRO membrane fouling may justify reduction of energy consumption for pretreatment. Given that the power requirement for pumping water from Essex station were not included in the parasitic load analysis, and that initial estimates of this power consumption showed it to be near the net power we computed, it is unclear whether this project is technically feasible at the site.

48

Works Cited
ABB. (2011). Motors and Generators. Retrieved from ABB.com: http://www.google.com/url?sa=t&rct=j&q=&esrc=s&source=web&cd=1&ved=0CCkQFjAA&url= http%3A%2F%2Fwww05.abb.com%2Fglobal%2Fscot%2Fscot234.nsf%2Fveritydisplay%2F7edf76 5650382c00c12578a8001acb6d%2F%24file%2Ftechnical_note_induction_vs_synchronous_how _to_select_ Alternate Hydro Energy Center. (2011). Standards/Manuals/Guidelines for Small Hydro Development. Dehli: Ministry of New and Renewable Energy, Government of India. Anagnostopoulos, J. S., & Papantonis, D. E. (2008). Flow Modeling and Runner Design Optimization in Turgo Water Turbines. International Journal of Engineering and Applied Sciences, 136-141. Andritz Hydro. (2013). Small-scale Hydropower Plants. Retrieved November 12, 2013, from Andritz: www.andritz.com/hydro/hy-small-hydropower-standard.htm Aziz, N. M., & Desai, V. R. (1991). An Experimental Study on the Effect of Some Design Parameters on Crossflow Turbine Efficiency. Clemson: Clemson University, Department of Civil Engineering. Boyle, G. ( 2004). Renewable Energy (2nd ed.). Oxford: Oxford University Press. British Stainless Steel Association. (2012). Selection of 316, 304, and 303 types of stainless steel for seawater application. Retrieved October 12, 2013, from British Stainless Steel Association: www.bssa.org.uk/topics.php?article=100 Brown, B. (2006). Consulting Engineer. Retrieved from Schneider-Electric.us: http://www.schneiderelectric.us/sites/us/en/customers/consulting-engineer/design-guide.page Canyon Hydro. (2013). Hydro Systems for Utilities and IPPs. Retrieved from CanyonHydro.com: http://www.canyonhydro.com/products/products.html Canyon Hydro Representative. (2013, October). Phone Conversation. (S. Clowser, Interviewer) CEC. (2008, 1 1). Renewables Portfolio Standard Eligibility Commission Guidebook. Retrieved 12 8, 2013, from energy.ca.gov: http://www.energy.ca.gov/2007publications/CEC-300-2007-006/CEC-3002007-006-ED3-CMF.PDF Copper Development Association. (1986). Meterials for Seawater pipeline Systems. CDA Publications. CPUC. (2011, 12 16). Virtual Net Metering. Retrieved 12 4, 2013, from cpuc.ca.gov: http://www.cpuc.ca.gov/PUC/energy/DistGen/vnm.htm Crewdsonem, V. E. (2013). Turbina Turgo. Retrieved December 2, 2013, from Energetika: http://mve.energetika.cz/primotlaketurbiny/turgo.htm

49

Dixon, S. L., & Hall, C. A. (2010). Fluid Mechanics and Thermodynamics of Turbomachinery. Burlington, MA, USA: Elsevier Inc. EPA. (2013, 11 14). Renewable Energy Production Incentives. Retrieved 12 4, 2013, from epa.gov: www.epa.gov/osw/hazard/wastemin/minimize/energyrec/rpsinc.htm European Small Hydro Association. (1998). Layman's Guidebook on How to Develop a Small Hydro Site. Brussels: Commision of European Communities. European Small Hydro Association. (2010). Energy Recovery in Existing Infrastructures with Small Hydropower Plants. Montcherand, Switzerland: Mhylab. European Small Hydropower Association. (2004). Guide on How to Develop a Small Hydropower Plant. Retrieved from Canyon Hydro: http://www.canyonhydro.com/images/Part_2_ESHA_Guide_on_how_to_develop_a_small_hydr opower_plant.pdf Francis, R. (2000, March 1). Pushing Belt Drive Efficiency. Retrieved December 3, 2013, from Machine Design: http://machinedesign.com/technologies/pushing-belt-drive-efficiency Free Flow Hydro. (2010). Hydroturbine-Crossflow-Ossberger. Retrieved November 8, 2013, from Free Flow Hydro: http://www.freeflowhydro.co.uk/13/24/CINK/CINK.html Fujihara, T., Imano, H., & Oshima, K. (1998). Development of Pump Turbine for Seawater PumpedStorage Power Plant. Hitachi Review, 199-202. Gilkes Hydropower. (2013). Turbine Selection. Retrieved October 18, 2013, from Gilkes Company: http://www.gilkes.com/Turbine-Selection HAP. (2011). Hydropower Advancement Project. Retrieved from Oak Ridge National Laboritory for the U.S. Deptartment of Energy: http://hydropower.ornl.gov/HAP/catalog.shtml Harvey, A. (1993). Micro-Hydro Design Manual. London: Intermediate Technology Publications. Hassan, A. M., & Malik, A. U. (1989). Corrosion Resistant Materials for Seawater RO Plants. Desalination, 157-170. Heng, S. (1992). Design of a 5 kW microhydro gernating set. Retrieved from University of Canterbury: http://ir.canterbury.ac.nz/handle/10092/6411 Hydrolink. (2009). Spiral Francis Turbine. Czech RePublic. Lako. (2010, 5 1). Hydropower. Retrieved 12 9, 2013, from Etsap: http://www.iea-etsap.org/web/etechds/pdf/e07-hydropower-gs-gct.pdf Loyer. (2013, 10 9). Certification as a renewable energy generation facility through CEC. (Cooper, Interviewer) CA, US. 50

Mann. (1999). High Energy Particle Impact Wear Resistance of Hard Coatings and Their Application In HyddroTurbines. Elsevier. Mathema, S. (2013). Electromechanical Aspects of Micro-Hydro and Small Hydro. Retrieved from Sari Energy: http://www.sarienergy.org/pagefiles/what_we_do/activities/afghan_capacity_building_program_aug_09/micro -hydro_0909/Presentations/English/Electro%20Mechanical%20Aspects%20of%20Micro%20Hydro%20&%2 0Small_Surendra%20Mathema_9th.pdf Meazza, G. (1996). Pelton Turbine. Retrieved October 22, 2013, from GM Hydro: http://www.gmhydro.it/EPelton/EPelton.htm Microhydro. (2013). Types of Waterwheels and Turbines. Retrieved December 17, 2013, from Microhydro Ireland: http://www.microhydro.ie/section/TypesofTurbines Morrow, S. J. (2010). Materials Selection for Seawater Pumps. Proceedings of the 26th International Pump Users Symposium (pp. 73-80). Lancaster: ITT Corporation. Murray. (2013, 11 29). FERC regulations for new energy generation facility projects. (Clowser, Interviewer) Nasir, B. A. (2013). Design of High Efficiency Pelton Turbine for Micro Hydropower Plant. International Journal of Electrical Engineering, 171-183. NiDI. (1995). Materials for saline water, desalination, and oilfield brine pumps. Nickel Development Institute. Norwol Power Systems. (2012, Aug 12). Single Phase vs. Three phase Generators. Retrieved from Norwol.com: http://www.norwall.com/blog/news-and-updates/single-phase-versus-threephase-generators/ OSSBERGER. (2013, 1 1). The Ossberger Turbine. Retrieved from www.ossberger.de: http://www.ossberger.de/cms/en/hydro/the-ossberger-turbine-for-asynchronous-andsynchronous-water-plants/ Paish, O. (2002). Small Hydro Power: Technology and Current Status. Renewable and Sustainable Energy Reviews, 537-556. PDH. (2012). Retrieved from PDH Engineering: http://electriciantraining.tpub.com/14177/css/14177_67.htm PG&E. (2013, 1 1). ReMAT Feed in Tariff (Senate Bill 32). Retrieved 12 5, 2013, from www.pge.com: http://www.pge.com/b2b/energysupply/wholesaleelectricsuppliersolicitation/ReMAT/

51

Photonics LTD. (2002). Power Factor. Retrieved from lmphotonics.com: http://www.lmphotonics.com/pwrfact.htm Raabe, J. (1985). Hydro Power: The Design, Use, and Function of Hydromechanical, Hydraulic, and Electrical Equipment. Dusselorf. Schultz, R. (1985). Introduction to Electric Power Engineering. New York: Harper & Row. Shahl, D. S. (2013). Induction Machines. Retrieved from http://www.google.com/url?sa=t&rct=j&q=&esrc=s&source=web&cd=5&ved=0CEkQFjAE&url=h ttp%3A%2F%2Fwww.uotechnology.edu.iq%2Fdepeee%2Flectures%2F3rd%2FElectrical%2FMachines%25202%2FV_IG.pdf&ei=X2uhUsnALMXzoAS m_IDIDg&usg=AFQjCNGOwQng1uN6kGHO3MArxljtvBwmPw&sig2 SWECO Norge AS. (2012). Cost base for small scale hydropower plants. Middelthunsgate: norwegian Water Resources and Energy Directorate. Taylor Engineering. (2013). Pipe Flow Tool. Retrieved December 14, 2013, from Taylor Engineering: http://www.taylor-engineering.com/track/index.htm Tcheslavski, D. G. (2008). Synchronous Machines. Retrieved from http://eng.uokerbala.edu.iq/lectures/electrical_engineering/Second_year/Machines/lecture%2 0%20-%20synchronous%20generators.pdf Todd, B. (1986). Nickel Containing Materials in Marine and Related Environments. Birmingham, UK: Nickel Development Institute. U.S. Army Corps of Engineers. (1979). Feasibility Studies for Small Scale Hydropower Addtions. Alexandria: U.S. Army Corps of Engineers. VDE. (2008, April 21). Positionspapier zur Wasserkraft. Retrieved December 16, 2013, from VDE: http://www.vde.com/de/fg/ETG/Arbeitsgebiete/V1/Aktuelles/Oeffentlich/Seiten/Wasserkraft.a spx Warnick, C. C. (1984). Hydropower Engineering. Englewood Cliffs: Prentice-Hall. Zia, O., Ghani, O. A., Wasif, S. T., & Hamid, Z. (2010). Design, Fabrication, and Installation of a MicroHydro Power Plant. Topi, Pakistan: GIK Institute of Engineering Sciences and Technology.

52

Appendix
The following pages contain the design input sheet, the output sheets for the impulse turbines, the pipe sizing and cost calculation tool inputs and outputs, the parasitic power loading provided by the other groups, and the final cost analysis tables calculated and presented, along with the figures.

53

GlobalInputs DecisionVariables Parameter Variable SaltwaterFreshwaterRatio R= PercentPermeate %p= OperationPressure P(kPa)= Pressure(psi) P(psi)= SystemParameters Variable Qg(MGD)=

Value

Notes 1.5 Varyasneeded 0.9445 Varyasneeded 1550 Frommembranegroup 224.812

Parameter GlobalFlow

Value

Notes 30 Fromclassdiscussion

CalculatedVariablesDoNotAlter Parameter Variable Value Totalfreshwaterflow Qtf(m3/s)= 0.544 Freshwaterpermeate Qfw(m3/s)= 0.514 Saltwaterflow Qsw(m3/s)= 0.771 FlushingFlow Qf(m3/s)= 0.030 Constants Parameter Variable DensityFreshWater (kq/m3)= DensitySaltWater (kq/m3)= DensityofBrackishWater (kq/m3)= Outputs Parameter Variable PowertoSaltwaterPressurising PSW(kW)= PowerGenerated(Pelton) Pgen(kW)= PowerGenerated(Francis) Pgen(kW)= PowerGenerated(Crossflow) Pgen(kW)= PowerGenerated(Turgo) Pgen(kW)=

Notes (22.8245MGD=1m3/s)

Value

Notes 998 (20C,1atm) 1030 (Beicher,RobertJ.PhysicsforSciandEngr) 1017.2 MixingEquation

Value

Notes 89.1 FromERDpage 644.2 FromPeltonDesignpage 671.3 FromCanyonHydroQuote 615.8 FromCrossflowDesignpage 622.7 FromTurgoDesignpage

PeltonWheelDesignCalculations SystemParameters Variable Value Notes Qs(mgd)= 17.6 Qf(mgd)= 12.4 = 0.9445 p(kPa)= 1502.2 (kg/m3)= 1017.2 g(m/s2)= 9.81 Q(m3/s)= 0.513677 ~11.75MGD@95%fluxandR=1.5 c(m/kPa)= 0.100213 (=1000/[rho*g]) PeltonWheelDesign Variable Value Notes Hnet(m)= 150.541 (N/m3)= 9978.732 Pth(kw)= 771.6503 n = 0.97 (typically0.960.99) (deg)= 165 (typically160170degrees) m= 0.95 (0.96isagoodvalue,HydroEngrbook) N(rpm)= 600 (desired) g= 0.95 i= 0.975 = 0.189031 (baseoffFluidsmechofturbomach.0.54.0) nj= 2 (baseonoperatingrangesforspecspeed) o= 1.00 (HydroEngineeringbook) Vj(m/s)= 52.71676 qj(m3/s)= 0.256839 dj(m)= 0.058162 (HydroEngineeringbook) u(m/s)= 25.22617 (HydroEngineeringbook) Dr(m)= 0.802974 (diameterofPelton@centerofjetimpact) 4.060506 (generalrule) Nb= 21 (Nasir) Bw(m)= 0.197752 (Nasir) Bl(m)= 0.174487 (Nasir) Bd(m)= 0.069795 (Nasir) sb(m)= 0.120125 Xjb(m)= 0.501859 Ft(N)= 27569.06 T(Nm)= 11068.62 (rad/s)= 62.83185 Po(kW)= 695.4619 Pn(kW)= 644.1716 s= 90.1% t= 83.5% Es(kWh/m3)= 0.37608

Parameter Seawatertotalflow: Freshwatertotalflow Fluxfraction: DesignPressure: Densityofdilutedseawater: Gravityconstant: DesignFlow(FWflux): Converisonfactor:

Input NetHead(Hh): SpecificWeightofwater: Theoreticalpoweroutput: NozzleLossCoef.: Bucketangle: Bucket(vane)lossfactor: Truerotationalspeed: Generatorlosscoef: Speedincreaserloss: DimensionlessSpec.Speed: NumberofJets: EfficiencyLossadj.from: Velocityofjet: Jetflow(each): Jetdiameter: Velocityofrunner(linear): Runnerdiameter: Runnerdiametercheck: Numberofbuckets: Bucketwidth: Bucketradiallength: Bucketdepth: Bucketspacing: Minimumjetclearance: TotalForce Torque: RadialSpeed: Poweroutput: Postgeneratorpower: TurbineEfficiency TotalEfficiency: Specificenergyextracted:

TurgoWheelDesignCalculations SystemParameters Variable Value Notes Qs(mgd)= 17.6 Qf(mgd)= 12.4 = 0.9445 p(kPa)= 1502.2 (kg/m3)= 1017.2 g(m/s2)= 9.81 Q(m3/s)= 0.51368 c(m/kPa)= 0.10021 (=1000/[rho*g]) TurgoRunnerDesign Variable Value Notes Hnet(m)= 150.541 (N/m3)= 9978.73 Pth(kw)= 771.65 n = 0.97 (typically0.960.99) 1(deg)= 20 2(deg)= 10 t(deg)= 150 (full,minusentry&exitfromhoriz) m= 0.95 (0.96isagood,HydrEngrbook,less4turgo) N(rpm)= 1200 (desired) g= 0.95 i= 1 = 0.37806 (baseoffFluidsmechofturbomach.0.54.0) nj= 2 (baseonoperatingrangesforspecspeed) nb= 20 (typicallyexperimentallydesignforsim.) Vj(m/s)= 52.7168 qj(m3/s)= 0.25684 dj(m)= 0.05816 (HydroEngineeringbook) u(m/s)= 23.7948 (HydroEngineeringbook) Dr(m)= 0.37871 (diameterofTurgo@centerofjetimpact) sb(m)= 0.05949 Ft(N)= 27545.2 T(Nm)= 5215.76 (rad/s)= 125.664 3.787059512 Po(kW)= 655.431 Pn(kW)= 622.66 s= 84.9% t = 80.7% Es(kWh/m3)= 0.35443

Parameter Seawatertotalflow: Freshwatertotalflow Fluxfraction: DesignPressure: Densityofdilutedseawater: Gravityconstant: DesignFlow(FWflux): Converisonfactor:

Input NetHead(Hh): SpecificWeightofwater: Theoreticalpoweroutput: NozzleLossCoef.: Bucketentryangle: Bucketexitangle: Bucketentryangle: Bucket(vane)lossfactor: Truerotationalspeed: Generatorlosscoef: Transmissionlosscoef.: DimensionlessSpec.Speed: NumberofJets: Numberofbuckets: Velocityofjet: Jetflow(each): Jetdiameter: Velocityofrunner(linear): Runnerdiameter: BucketSpacing: TotalForce Torque: RadialSpeed: Poweroutput: Postgeneratorpower: TurbineEfficiency TotalEfficiency: Specificenergyextracted:

Parameter Seawatertotalflow: Freshwatertotalflow Fluxfraction: DesignPressure: Densityofdilutedseawater: Gravityconstant: DesignFlow(FWflux): Converisonfactor:

CossflowTurbineDesignCalculations SystemParameters Variable Value Notes Qs(mgd)= 17.6 Qf(mgd)= 12.4 = 0.9445 p(kPa)= 1502.2 (kg/m3)= 1017.2 g(m/s2)= 9.81 Q(m3/s)= 0.5136771 c(m/kPa)= 0.1002131 (=1000/[rho*g]) CrossflowTurbineDesign Variable Value Hnet(m)= 150.54105 Hnet(ft)= 493.92518 (N/m3)= 9978.732 Pth(kw)= 771.65027 Q(cfs)= 18.140507 s= 84% D1(in)= 23.5 D2/D1= 0.67 D2(in)= 15.745 1()= 22 2()= 105.00 ()= 90 nB= 20 rs= 0.7 dp(in)= 18 Vp(ft/s)= 10.265427 vj(ft/min)= 10701.01 ur(ft/min)= 4960.9018 D2p(in)= 15.563443 x = 0.6268069 s(in)= 6.9140603 1()= 38.940029 b(in)= 3.4037724 ()= 73.420365 Rr(rpm)= 1739.3552 Rp(rpm)= 1217.5486 Rg(rpm)= 1200 rgb= 0.99 nw(in)= 6.4625227 rl(in)= 14.729963 rwa(in)= 3.8305 rw/sb= 1.0376902 Es(kWh/m3)= 0.3505154

Parameter NetHead(Hh): NetHeadinfeet(H) SpecificWeightofwater: Theoreticalpoweroutput: DesignFlowinenglishunits: ExpectedEfficiency: OuterDiameter: DiameterRatio: Innerdiameter: Angleofattack(Alpha1) Bladeexitangle(Beta2) Nozzleentryarc(Lambda) Numberofblades(nB) Speedratio(Azizsays0.600.70) InflowPipediameter Pipevelocity Speedofwaterjet Runnertangentialvelodity Innerdiameterdesigncheck(Zia) AspectratioB/D1(.33.50pref.) Nozzlethroatwidth Bladeinletangle Radiusofbladecurvature Bladecentralangle Turbinerunawayspeed Peakefficiencyspeed GeneratorSpeed Gear/beltratio Nozzlewidth Runnerlength(Aziz) Runnerbladeradiusofcurvature: Runnerwidth/bladespacingratio SpecificEnergyExtracted

Notes

(reachingmaxpossible) (0.60.7best,Zia,Aziz) (Aziz,upto24fromZiaet.Al.) (90isideal,Aziz) (experimentallydet.Upto37) (0.54fromAziz&Desai,0.67Zia) (justforconsistencyfromERD)

(Ziaet.Al) (Ziaet.al) (Aziz&Desai) (Aziz&Desai)

(<3forbelt,=1.00fordirect)

(Ziaet.Al) (1.03isoptimum,Aziz) [Po*1hr/Q(m3/s)*3600sec]

OutputfromTurbine GeneratorLosscoefficient FinalOutput Belttension(forabeltdrive) FinalEfficiency MetricSpecificSpeed SpeedofJet Speedofrunner BladeSpacing InnerDiameter OuterDiameter SpecificEnergyExtracted

FinalEstimatedOutputs Po(kW)= 648.18623 g= 0.95 Pf(kW)= 615.77692 Tb(lb)= 3829.366 t = 79.8% Ns= 57.937863 vj(m/s)= 54.358478 vr(m/s)= 37.502503 sb(m)= 0.0937563 D2(m)= 0.5968709 D1(m)= 0.3999035 Es(kWh/m3)= 0.3505154

Finalvaluesestimatedfromtheory andcalculationpageusingAziz turbinesources

Clear and Apply User Settings

Auto-select Optimized Pipe Size

Inputs CRITICAL Select Pipe Size or Pipe "AUTO" Size rounded GPM 21000 21000 Flow Speed Limit to Prevent Noise? Too fast? NO NO Flow Speed Limit to Prevent Erosion? Too fast? NO NO Desired Insulation Wall Thickness

Total system flow

21000 GPM

Add or delete a row above current cursor position

Segment Grouping

Flow rate

Insulation Thickness

Straight Pipe Number of Length Elbows

Tees with flow through: Straight Branch Circuit Setter Silent Check

Number of Valves Ball or Butterfly Suction Diffuser Flow Limiting WyeStrainer Swing Check

Other: Chiller/ Control Valve Boiler/ Coil/ HX/ P etc.

Pipe Segment Description 1

GPM 21000 21000

Inches 30 30

Inches 30 30

ON ON

ON ON

Inches Title 24 Title 24

0 0

Feet 90 45 800 1 685 1 Outputs

Psi

Feet

1 Totals

Segment Head

Water Const/ Pump Friction volume in Variable Motor Rate pipe multiplier Power

Total Power

ELL TEES SK K MATERIALS FAC FACT COST: TOR OR PIPING

MATERIALS COST: ELLS

MAT ERIA LS COS T: TEES

Feet 8.6 5.7 14.3 5 19.3

psi 3.7 2.5 6.21

Feet/ 100 ft 0.8 0.8

Gal 27,927 23,913 51,840

therms kW kW /hr 45 LINE insulation 0.3 14.08 16.38 0.00 0.12 0.05 $ 0.3 9.30 10.81 0.00 0.12 0.05 $ without Safety Factor with Safety Factor

$ $

90 45 3,000.00 $ 3,000.00 $ -

MATERIALS COST: VALVES Circui Swin Flow t g Ball/ Limitin Sette Chec Butter g LINE r Silent Check k fly Valve insulation LINE ## ## $ 23,650.00 ## $ $ $ $ ## ## $ ## $ $ $ $ -

LABOR COST: PIPING

LABOR COST: TEES

LABOR COST: VALVES Flow Ball/ Limiti Butter ng fly Valve $$$$-

First Cost

Lifecycle + First Cost

$ $ $

Labor 136,475 $ 115,931 $ 3,151 $ $

Total 590,512 $ 911,535 479,267 $ 691,253 1,118,535 ######## 1,069,779

Summaryofpowerdrawsasprovidedbyothergroups

LoadsinkWasprovided Fw2 SaltWater ERD generated net parasitic 132 99 155 671 285 386

LoadsinkWasprovided Fw1 SaltWater ERD generated net parasitic 187 99 155 671 230 441

ParasiticLoad FreshWater2 132 SaltWater 99 ERD 155 NetPower 239 (forPieCharts)

ParasiticLoad FreshWater1 187 SaltWater 99 ERD 155 NetPower 210 (forPieCharts)

Totalsforyourcomponent Group FW1 FW2 Mem SW Tur ERD Total(FW1) Total(FW2) O&M($/year) AnnualEnergy Loss(kWh/yr) $432,133.00 $3,533,034.00 2,168,835 $324,884.00 $7,885,064.00 1,391,635 $440,000.00 $5,905,751.00 0 $124,976.00 $16,785,058.00 950,000 $71,880.00 $2,070,000.00 0 $50,000.00 $3,456,000.00 1,320,000 $1,118,989.00 $31,749,843.00 4,438,835 $1,011,740.00 $36,101,873.00 3,661,635.00 ParasiticLoad(FW1) 441 kW ParasiticLoad(FW2) 386 kW Capital($)

hrsinyear generation netpower netenergy PPA revenue rounded

8760 671 230.0 2014800 0.09375 $188,887.50 $189,000.00

hrs kW kW kWh $/kWh

hrsinyear generation netpower netenergy PPA revenue rounded

8760 671 285.3 2499578.4 0.09375 $234,335.48 $235,000.00

hrs kW kW kWh $/kWh

Costsforthesystembasedonothergroupsfindingsaslastreportedtous

OperationsandMaintenance FreshWater1 $432,133.00 CosttoDeliverFW $ 1,299,960 Membrane 440,000 Seawater $124,976.00 Turbine $71,880.00 ERD $50,000.00 Total $2,418,949.00 FreshWater2 CosttoDeliverFW Membrane Seawater Turbine ERD Total $324,884.00 $ 1,299,960 440,000 $124,976.00 $71,880.00 $50,000.00 $2,311,700.00

CapitalCost FreshWater1 $3,533,034.00 Membrane $5,905,751.00 Seawater $16,785,058.00 Turbine $1,050,000.00 ERD $3,456,000.00 Total $30,729,843.00 FreshWater2 $7,885,064.00 Membrane $5,905,751.00 Seawater $16,785,058.00 Turbine $1,050,000.00 ERD $3,456,000.00 $35,081,873.00 Total

Summaryofproducedcharts

Net Power 32%

Fresh Water1 29% Salt Water 15%

ERD 24%

Net Power 38%

Fresh Water2 21% Salt Water 16% ERD 25%

Seawater Turbine 5% 3%

ERD 2%

FreshWate r1 18%

Membrane 18%

Costto DeliverFW 54%

Turbine Seawater 3%
6% Membrane 19%

ERD 2%

FreshWate r2 14%

Costto DeliverFW 56%

Turbine 3%

ERD 11%

FreshWate r1 12%

Membrane 19% Seawater 55%

Turbine 3%

ERD 10%

FreshWate r2 22% Membrane 17%

Seawater 48%

You might also like