Case Details

A dozen Koi fish killed
Bushkill Township, PA (US)

Date: May 2004
Disposition: Open

Suspect(s) Unknown - We need your help!

Case ID: 2446
Classification: Poisoning
Animal: marine animal (pet)
View more cases in PA (US)
Reward: $1000.00
Login to Watch this Case

Steven Sauerzopf, or Bushkill Township, PA is offering a $1,000 reward for information leading to the arrest and conviction of the person who killed a dozen of his prized fish in his front yard.

Sauerzopf said his wife, Linda, discovered the dying fish on Saturday. After draining the small pond, she found a sponge with a piece of nylon string attached to it. The sponge was sent to a lab for testing by Bushkill police.

Linda managed to save about 18 smaller koi. The dead fish had grown from just a few inches to about two feet.

"You can't buy fish that long. They would cost anywhere from $500 to $1,500 a piece," Sauerzopf said.

The Sauerzopfs own L&S Transportation, a small bus service company that they run out of their home. The Sauerzopfs have had a longstanding dispute with their neighbors, who say the business does not belong in the neighborhood.

Steven Sauerzopf said he suspects that whoever poisoned his fish might also have poisoned shrubs in his front yard.  Pellets have also been shot at Sauerzopf's property.

Anyone with any information is asked to contact Steven, care of Bushkill police.

If you have information on this case, please contact:
Steven Sauerzopf C/O Bushkill Police
(570) 775-7374

References

www.nepanews.com
www.poconorecord.com

« Back to Search Results

Note: Classifications and other fields should not be used to determine what specific charges the suspect is facing or was convicted of - they are for research and statistical purposes only. The case report and subsequent updates outline the specific charges. Charges referenced in the original case report may be modified throughout the course of the investigation or trial, so case updates, when available, should always be considered the most accurate reflection of charges.

For more information regarding classifications and usage of this database, please visit the database notes and disclaimer.



Send this page to a friend
© Copyright 2001-2006 Pet-Abuse.Com. All rights reserved. Site Map ¤ Disclaimer ¤ Privacy Policy